INVITED REVIEW
Proteinase-activated receptors: novel mechanisms of signaling by serine proteases

Olivier Déry, Carlos U. Corvera, Martin Steinhoff, and Nigel W. Bunnett

Departments of Surgery and Physiology, University of California, San Francisco, California 94143-0660

    ABSTRACT
Top
Abstract
Introduction
Conclusions
References

Although serine proteases are usually considered to act principally as degradative enzymes, certain proteases are signaling molecules that specifically regulate cells by cleaving and triggering members of a new family of proteinase-activated receptors (PARs). There are three members of this family, PAR-1 and PAR-3, which are receptors for thrombin, and PAR-2, a receptor for trypsin and mast cell tryptase. Proteases cleave within the extracellular NH2-terminus of their receptors to expose a new NH2-terminus. Specific residues within this tethered ligand domain interact with extracellular domains of the cleaved receptor, resulting in activation. In common with many G protein-coupled receptors, PARs couple to multiple G proteins and thereby activate many parallel mechanisms of signal transduction. PARs are expressed in multiple tissues by a wide variety of cells, where they are involved in several pathophysiological processes, including growth and development, mitogenesis, and inflammation. Because the cleaved receptor is physically coupled to its agonist, efficient mechanisms exist to terminate signaling and prevent uncontrolled stimulation. These include cleavage of the tethered ligand, receptor phosphorylation and uncoupling from G proteins, and endocytosis and lysosomal degradation of activated receptors.

G protein-coupled receptors; thrombin; trypsin

    INTRODUCTION
Top
Abstract
Introduction
Conclusions
References

SERINE PROTEASES COMPRISE a large family of enzymes that are characterized by a uniquely reactive Ser side chain. They are ubiquitous in prokaryotes and eukaryotes and serve important and diverse biological functions. These include hemostasis, fibrinolysis, complement formation, and the digestion of dietary proteins. Accumulating evidence indicates that certain serine (Ser) proteases, which have been traditionally considered to participate principally in the degradation of extracellular proteins, are also signaling molecules that regulate multiple cellular functions by activating specific receptors. At least three types of protease receptors have been identified. The receptors for coagulation factor Xa and urokinase are examples of single transmembrane domain and glycosylphosphatidylinositol-linked receptors, respectively, which do not require proteolytic cleavage for activation. The third emerging family of protease receptors, which is the topic of this review, are G protein-coupled receptors (GPCRs) that are activated by proteolysis (Fig. 1). There are three known members of this receptor family of proteinase-activated receptors (PARs): PAR-1 and PAR-3, which are cleaved and activated by thrombin, and PAR-2, a receptor for trypsinlike enzymes.


View larger version (29K):
[in this window]
[in a new window]
 
Fig. 1.   Mechanism of G protein-coupled receptor (GPCR) activation by a reversible binding of a soluble ligand, such as a neuropeptide, and irreversible cleavage by a protease.

We discuss the evidence that certain proteases function as signaling molecules to specifically regulate target cells by cleaving PARs. This topic is of interest since it defines novel functions for proteases, which are traditionally viewed as degradative enzymes. Furthermore, a comparison of the mechanisms of activation, signaling, and regulation by PARs and receptors for other signaling molecules, in particular for neuropeptides, provides new insights into the biology of GPCRs in general. Consequently, where appropriate, we compare signaling by Ser proteases and PARs with signaling by neuropeptides and their receptors. Because this review is not a complete survey of all the literature on PARs, the reader may wish to refer to other recent reviews (21, 55, 61).

    TURNING ON THE SIGNAL: MECHANISMS OF PAR ACTIVATION

Although agonists of GPCRs are extremely diverse, ranging from photons to Ser proteases, receptor activation involves several steps that are common for most receptors (Fig. 1). The first step involves interaction between the ligand and its receptor. Agonists of most known GPCRs are small, hydrophilic molecules, such as catecholamines or peptides, which are soluble in the extracellular fluid. These agonists bind to extracellular and transmembrane domains of their receptors at the cell surface. However, Ser proteases activate PARs by a unique process that involves recognition of the receptor by the enzyme, cleavage of the receptor at a specific enzymatic site within the extracellular NH2-terminus, and finally exposure of a new NH2-terminus that acts as a tethered ligand, which binds and activates the cleaved receptor molecule. Specific residues within the tethered ligand domain, which usually comprises six or more amino acids, interact with specific extracellular and transmembrane domains of the cleaved receptor. Therefore, PARs may be considered to be specialized peptide receptors: ones in which the ligand is physically part of the cleaved receptor molecule. The second step of GPCR activation is a change in the conformation of the receptor, although the nature of this change is not understood. Finally, receptors in an active conformation interact with heterotrimeric G proteins in the plasma membrane, which transduce the signal. This section discusses these mechanisms of activation for PARs.

Activation of PAR-1

PAR-1 is the most extensively studied receptor of this family, and the mechanism of activation is well defined (Fig. 2). The extracellular, NH2-terminus of human PAR-1 contains a putative cleavage site for thrombin (LDPR41down-arrow S42FLLRN) followed by a sequence of charged residues (D51KYEPF56) (160, 161). The charged domain interacts with an anion binding site on thrombin, probably inducing conformational changes in the receptor to accommodate its cleavage site into the thrombin catalytic subsites and to promote efficient receptor hydrolysis. The negatively charged region of PAR-1 resembles a domain in the COOH-tail of the leech anticoagulant hirudin, which inhibits the thrombin by binding its anion site. The importance of this domain is indicated by the finding that its deletion leads to a loss of activation of the receptor by thrombin, whereas substitution of this region by the corresponding domain of hirudin allows a full recovery of activity (161).


View larger version (29K):
[in this window]
[in a new window]
 
Fig. 2.   A: protein structure of protein-activated receptor (PAR)-1, PAR-2, and PAR-3. Amino acid sequences in NH2-terminus and second extracellular loop that are important for receptor activation are shown. Boxed residues indicate tethered ligand domains (PAR-1, PAR-2, and PAR-3) and anion binding sites (PAR-1 and PAR-3). Arrows indicate cleavage sites. Bold residues in second extracellular loop are conserved. # Intron/exon border. * Glycosylation site. B: genomic organization and chromosomal localization of PAR-1 and PAR-2. Both genes consist of 2 exons and 1 large intron. Exon 1 encodes NH2-terminal domains proximal to cleavage sites, and exon 2 encodes the rest of the receptors. Both receptors are localized within 100 kb on chromosome 5q13.

Several strategies have been used to demonstrate that thrombin cleaves PAR-1. Analysis of platelets by flow cytometry using region-specific PAR-1 antibodies provides physical proof for receptor cleavage. Thus thrombin treatment reduces cell surface immunoreactivity for a polyclonal antibody directed against the 34-52 region of the receptor (which spans the cleavage site), whereas binding of an antibody directed against the 83-94 region (which is COOH-terminal to the cleavage site) is not decreased by thrombin (114). Analysis of epitope-tagged PAR-1 by Western blotting provides further evidence for cleavage. Mature PAR-1 has a molecular mass of 68-80 kDa, which is reduced to 36-40 kDa by deglycosylation (158). The electrophoretic mobility of PAR-1 in membrane proteins from transfected cells is increased after activation by thrombin, indicating a reduction of molecular mass by proteolytic cleavage.

The mechanism of PAR-1 activation has also been investigated by genetic analysis, involving generation of receptor chimeras and point mutants. An epitope-tagged protein, corresponding to the NH2-terminus of PAR-1 fused with the transmembrane domain of CD8, has been used to evaluate the determinants of efficient cleavage by thrombin (73). The rate of thrombin cleavage, assessed by loss of receptor immunoreactivity from the cell surface, is similar for this chimera and the wild-type receptor. The rate of cleavage is unaffected by substitution of the anionic site by the corresponding domain of hirudin. Together, these results indicate that the extracellular NH2-terminus of PAR-1, which includes the cleavage domain (LDPR41down-arrow S42FLLRN) and the negatively charged domain (D51KYEPF56), is sufficient for efficient cleavage by thrombin. Furthermore, thrombin efficiently cleaves a soluble peptide corresponding to the NH2-terminus of PAR-1 (Km = 15-30 µM, kcat = 50 s-1), generating a product with the sequence SFLLRN, providing biochemical proof for cleavage at this site (73). Analysis of another fusion protein formed between glutathione-S-transferase and the 25-97 sequence of PAR-1 demonstrates that alpha -thrombin cleaves PAR-1 between Arg41 and Ser42, whereas gamma -thrombin, which lacks the anion binding site, is 100-fold less potent (19). The importance of this site for activation is also illustrated by the finding that a Ser42 to Pro mutation abolishes cleavage and activation by thrombin (160, 161).

Thrombin cleaves PAR-1 to expose a new NH2-terminus (S42FLLRN), which binds to the activation site of the receptor. Indeed, synthetic peptides that correspond to this tethered ligand domain activate PAR-1 without cleavage of the receptor (135, 160). Analogs of the tethered ligand are of great interest since they form a starting point for generating specific receptor agonists or antagonists and may thus be used to probe the function of PARs without the use of proteases, which may cleave other proteins and thus exert effects by several mechanisms. Analysis of analogs of SFLLRN in which individual residues are substituted, together with site-directed mutagenesis of the tethered ligand, indicates that critical residues within this domain include Phe2, Leu4, and Arg5 (135). Unfortunately, peptides corresponding to the tethered ligand domain are relatively weak agonists compared with proteases. Differences in potency are probably due to the inefficient presentation of these soluble peptides to the binding domains of the receptor, compared with the tethered peptide. In addition, these peptides are readily inactivated by proteolysis (53).

The interaction between the tethered ligand and PAR-1 is critical for transmembrane signaling of this receptor. This interaction has been investigated by analysis of chimeras of human and Xenopus PAR-1 (50, 105). Xenopus PAR-1 has a tethered ligand with the sequence TFRIFD. Bioassays of the peptides corresponding to the tethered ligands of human and Xenopus PAR-1, or substituted analogs, indicate that these peptides are specific for their parent receptors (50). Replacement of the extracellular face of Xenopus PAR-1 with corresponding domains of human PAR-1 switches selectivity from the Xenopus-derived to human-derived peptides (Fig. 3). Similarly, substitution of the extracellular NH2-terminus and the second extracellular loop of the Xenopus receptor with the corresponding domains of the human receptor also confers the chimera with selectivity for peptides corresponding to the human tethered ligand. These studies indicate that these domains are sufficient to confer Xenopus PAR-1 with selectivity for peptides corresponding to the human tethered ligand. The docking interactions have been investigated in more detail by analysis of receptor chimeras with point mutations. Substitution of two residues of Xenopus PAR-1 with corresponding residues of the human receptor (Asn87right-arrowPhe in the extracellular NH2-terminus, and Leu260right-arrowGlu in the second extracellular loop) is sufficient to confer selectivity for human peptides (105). This finding indicates that these residues in the extracellular NH2-terminus and extracellular loop 2 are critical for interaction with the tethered ligand domain. Further analysis of mutated receptors and mutated agonist peptides suggests that Glu260 of extracellular loop 2 interacts with Arg5 of the tethered ligand SFLLRN (105). Finally, substitution of eight residues from the second extracellular loop of the Xenopus receptor to human PAR-1 results in elevated 45Ca2+ release from oocytes, suggesting that the receptor is constitutively active, even in the absence of thrombin (106). Therefore, an alteration in the conformation of the extracellular loops of PAR-1 is sufficient to transduce a signal across the plasma membrane to G proteins that is similar to that observed after receptor cleavage or interaction with peptides corresponding to the tethered ligand domain. This observation also underscores the importance of extracellular loop 2 for receptor activation and regulation (106).


View larger version (29K):
[in this window]
[in a new window]
 
Fig. 3.   Receptor domains mediating the specificity of activation of the thrombin receptor (PAR-1). Wild-type and chimeric receptors were expressed in Xenopus oocytes, and EC50 values were measured for the human-selective (SFLLRN) and Xenopus-selective (TFRIFD) peptide agonists. {Reprinted by permission from Nature [Gerszten et al. (50)] copyright 1994 Macmillan Magazines Ltd.}

The principal mechanism of PAR-1 activation is intramolecular: interaction of a tethered ligand with the receptor molecule to which it is attached. However, there is also evidence for intermolecular signaling (28). Thus coexpression of nonactivatable PAR-1 mutant (with a deleted NH2-terminus) and a chimera composed of the NH2-terminus of PAR-1 with the transmembrane domain of CD8, generates cells that respond to thrombin, suggesting that the tethered ligand of the chimeric receptor is able to activate the mutant PAR-1. This result supports the hypothesis of an intermolecular interaction between cleaved and uncleaved PAR-1, although the intramolecular mechanism of activation remains predominant (28). Whether intramolecular signaling occurs between different PARs is unknown. Indeed, synthetic peptides derived from the tethered ligand of PAR-1 can activate both PAR-1 and PAR-2, although PAR-1 is selective for its own agonist peptides (13).

Activation of PAR-2

The second member of this receptor family, PAR-2, was identified by screening a mouse genomic library by PCR using degenerate primers to the second and sixth transmembrane domains of the neurokinin 2 receptor (Fig. 2) (116, 117). A clone was identified that encoded a protein with the typical characteristics of a GPCR with ~30% amino acid identity to human PAR-1. The long extracellular NH2-terminus of mouse PAR-2 contains a putative trypsin cleavage site SKGR34down-arrow S35LIGR. Subsequently, PAR-2 has been cloned in human and rat (18, 115, 132).

Several lines of evidence indicate that trypsin cleaves PAR-2, exposing a tethered ligand that binds and activates the cleaved receptor. Mutation of Ser34 to Pro prevents cleavage and activation of PAR-2 by trypsin (116). Exposure of transfected cells to trypsin results in loss of immunoreactivity to an antibody against an NH2-terminal epitope, which indicates that trypsin cleaves intact PAR-2 at the plasma membrane (17). Furthermore, analysis of cleavage of synthetic peptides corresponding to a potential trypsin site indicates that trypsin hydrolyses the Arg34-Ser35 bond (98). Synthetic peptides corresponding to the tethered ligand domain (SLIGRL in mouse, SLIGKV in human) activate PAR-2 in transfected cell lines without the need for receptor cleavage (18, 115-117). These peptides also mimic the effects of trypsin in cell lines that naturally express PAR-2, such as epithelial cells (enterocytes, keratinocytes), endothelial cells, myocytes, T cell lines, neutrophils, and various tumor cell lines (1, 18, 35, 66, 70, 83, 92, 96, 132, 133). Critical residues within the tethered ligand have been identified by assaying peptide analogs for biological activity (13, 63). Peptides of only five residues retain some activity, and activity is slightly enhanced by amidation of the COOH-terminus. Residues Leu2 and Arg5 are of critical importance. Together, these results suggest that proteases activate PAR-1 and PAR-2 in a similar manner, involving receptor cleavage to expose a tethered ligand. However, in contrast to PAR-1, where thrombin interacts with the receptor at the hirudin-like site, there is no evidence that trypsin interacts with PAR-2 at sites other than the cleavage site. It remains a possibility that other proteases that activate PAR-2 also interact with the receptor at additional sites.

The interactions between tethered ligands and the cleaved receptors have been further defined by examining the specificity with which peptides activate different PARs and by studying chimeric receptors derived from PAR-1 and PAR-2 (89). Cells expressing PAR-2 respond to both PAR-2- and PAR-1-derived peptides, whereas PAR-1 only responds to its own peptide (13). Analysis of chimeras between the human PAR-1 and PAR-2 corroborates the results obtained with chimeras of the human and Xenopus PAR-1 (50, 105). Substitution of the extracellular face of PAR-2 (NH2-terminus and the three extracellular loops) with the corresponding regions of PAR-1 yields a chimera with PAR-2-like selectivity for agonists (89). Substitution of individual extracellular domains again reveals the importance of the second extracellular loop in agonist activation.

Activation of PAR-3

The molecular cloning of PAR-2 suggested the existence of an extended family of PARs and provided the impetus for attempts to identify other receptors of this type (36). In some systems, peptides corresponding to the tethered ligands of PAR-1 are considerably less efficacious than thrombin, suggesting the existence of other thrombin receptors (80, 88, 90, 146). This suspicion was strengthened by the finding that platelets derived from mice, in which the PAR-1 gene was deleted by homologous recombination, respond to thrombin but not PAR-1 peptides (34). The existence of a second thrombin receptor, designated PAR-3, has now been confirmed by molecular cloning, in which degenerate primers to various domains of PAR-1 and PAR-2 were used to screen RNA from rat platelets by PCR (71). A clone was identified with ~28% amino acid sequence similarity to both human PAR-1 and PAR-2 (Fig. 2). There are several structural and functional similarities between PAR-1 and PAR-3. In common with PAR-1, PAR-3 contains a hirudin-like site (FEEFP in PAR-3) that interacts with thrombin (71, 160). Thus gamma -thrombin, which is defective in this binding site, is 100 times less potent than alpha -thrombin in activating PAR-3, and Ala substitutions within the hirudin site of PAR-3 attenuate activation of PAR-3 by thrombin. Thrombin cleaves PAR-3 at LPIK38down-arrow T39FRG, and mutation of the cleavage site to one that would be resistant to thrombin prevents activation. Cleavage by thrombin exposes a new NH2-terminus (TFRGAP) that may interact with the receptor as a tethered ligand. However, in marked contrast to PAR-1 and PAR-2, synthetic peptides corresponding to this putative tethered ligand do not activate PAR-3. Presently, there is no explanation for this difference, although differences in affinity, steric hindrances, and the possibility that cleavage releases conformation of the receptor constrained by the uncleaved NH2-terminus region may explain these unexpected results.

Activation by Other Proteases

Among the studied proteases, thrombin cleaves and activates PAR-1 and PAR-3 with the highest efficiency, and pancreatic trypsin is the most potent agonist of PAR-2. However, other proteases are also capable of cleaving and activating these receptors if they too cleave in a manner that exposes a tethered ligand domain (Fig. 4). Besides thrombin, trypsin, plasmin, granzyme A, and cathepsin G also activate PAR-1 by cleavage at the thrombin site (99, 147, 148, 158). In addition to pancreatic trypsin, other enzymes with trypsinlike selectivity cleave and activate PAR-2, notably mast cell tryptase and coagulation factor Xa (35, 98, and S. Böhm, C. Corvera, L. Khitin, W. Kong, G. Caughey, D. Payan, and N. Bunnett, unpublished observations). Thrombin is the principal activator of PAR-3, but factor Xa, trypsin, elastase, and chymotrypsin weakly stimulate 45Ca2+ efflux from Xenopus oocytes expressing PAR-3 (71).


View larger version (10K):
[in this window]
[in a new window]
 
Fig. 4.   Principal protease cleavage sites for PAR-1, PAR-2, and PAR-3. Cleavage will activate the receptors if it exposes the tethered ligand (bold arrow) but may inactivate the receptors if the tethered ligand is destroyed or removed (fainter arrows). CG, cathepsin; PR3, proteinase 3; HLE, human leukocyte elastase. (Data from Refs. 98, 99, 128.)

The physiological relevance of PAR activation by different proteases is uncertain. Clearly, these proteases may be the principal activators in certain tissues, particularly in the case of PAR-2. PAR-2 is highly expressed by enterocytes, where it may be activated by trypsin in the intestinal lumen (83). However, it is also expressed in tissues that are not exposed to trypsin, including intestinal smooth muscle and the skin. In these locations, mast cell tryptase and other unidentified proteases may cleave and trigger PAR-2 (35, 98, and S. Böhm, C. Corvera, L. Khitin, W. Kong, G. Caughey, D. Payan, and N. Bunnett, unpublished observations).

Deactivation by Other Proteases

Some proteases cleave PARs at sites other than the activating site to generate a receptor that is unresponsive to subsequent proteolytic activation (Fig. 4). Thus, in addition to the Arg41down-arrow Ser42 site, cathepsin G cleaves PAR-1 at Phe43down-arrow Leu44 and Phe55down-arrow Trp56, removing the tethered ligand and rendering the receptor unresponsive to thrombin (99). The neutrophil enzymes elastase, proteinase 3, and cathepsin G also cleave PAR-1 on platelets and endothelial cells downstream of the thrombin cleavage site, indicating that these enzymes can inhibit the biological actions of thrombin (128). Similarly, tryptase also cleaves PAR-2 at the Lys41down-arrow Val42 site, in addition to the activating cleavage at the site Arg34down-arrow Ser35 (98). However, the activating cleavage is probably favored, since tryptase stimulates cells expressing PAR-2. Hydrolysis of PARs at sites that result in removal of the tethered ligand may provide a mechanism for downregulating these receptors and may thus provide an additional regulatory mechanism.

Cleavage may also serve to terminate signaling of an activated PAR. In HEL megakaryoblast erythroleukemia cells, thrombin cleaves PAR-1 and stimulates a transient increase in intracellular Ca2+ concentration ([Ca2+]i) (29). In contrast, in insect SF9 cells expressing PAR-1, thrombin induces a sustained increase in [Ca2+]i, which slowly desensitizes. Addition of thermolysin to thrombin-stimulated SF9 cells attenuates this otherwise sustained response. Thermolysin may cleave PAR-1 within the tethered ligand domain (SFLLR) at Phe43-Leu44 and Leu44-Leu45, suggesting that thermolysin terminates signaling by inactivating the tethered ligand. Such inactivation may represent a novel mechanism of regulation, by which extracellular or cell surface proteases terminate signaling of PARs by cleaving tethered ligands.

Comparison Between Activation of PARs and Neuropeptide Receptors

A comparison of mechanisms of activation of receptors for proteases and peptides provides insights into signaling mechanisms and into the biology of GPCRs in general. In some respects, PARs can be considered as specialized peptide receptors, but ones in which the peptide ligand is physically part of the receptor molecule.

Proteolysis is important for the initiation of signaling by peptides and proteases. Postsecretory processing of peptide hormones and neurotransmitters is required to generate biologically active forms. Most peptides are synthesized as large, inactive precursors that are processed within cells by a family of prohormone convertases to the biologically active, secreted molecules. For example, posttranslational processing of the preprotachykinins generates the biologically active peptides substance P, neurokinin A, and neurokinin B (107). Some proteases, exemplified by angiotensin-converting enzyme, convert peptides to their principal biological forms in the extracellular fluid and thereby initiate signaling.

Despite their unique mechanism of activation, the observation that several different proteases cleave and activate the PARs with differing affinities is reminiscent of the ability of several neuropeptides to activate a particular receptor. Thus all three of the principal tachykinin peptides, substance P, neurokinin A, and neurokinin B, interact with the three main neurokinin receptors (NK1-R, NK2-R, and NK3-R), albeit with graded affinity (120). However, the physiological relevance of the ability of multiple agonists, whether proteases or peptides, to activate a single receptor is unknown.

There are also similarities between neuropeptide receptors and PARs in the second step of activation: interaction of the ligand with the receptor. Specific residues within the tethered ligand domain of PAR-1 interact with extracellular domains of the cleaved receptor. Similarly, residues in the COOH-terminus of substance P interact with extracellular domains of the NK1-R. However, in general, far more is known about the nature of the interaction between neuropeptides and their receptors than about interaction between the tethered ligand of PAR-1 and the cleaved receptor. More information is available about neuropeptide-receptor interaction, because multiple different methods are available to analyze such interactions. Thus the domains of the NK1-R that interact with peptide agonists and with nonpeptide antagonists have been defined by domain-swapping between different neurokinin receptors, by point mutation, including generation of receptors with metal binding domains, and by cross-linking experiments with photoactivatable agonists and subsequent biochemical analysis of receptor fragments that interact with these agonists (20, 43, 51, 52). This last approach indicates that the third and eighth positions of substance P interact with the extracellular tail (residues 1-21) and the second extracellular loop (residues 173-183), respectively, of the NK1-R (20). The general consensus of these approaches is that natural peptide agonists bind to residues scattered throughout the extracellular regions of the NK1-R, whereas nonpeptide antagonists interact with residues in the transmembrane domain. Application of similar approaches is expected to provide further information about the molecular mechanisms of PAR activation.

Although neuropeptide receptors and PARs couple to the same G proteins and activate the similar signaling cascades, there are also distinct differences, in particular, those related to the mechanism by which cells respond to variable concentrations of neuropeptides and proteases. Neuropeptides elicit graded cellular responses by graded receptor occupancy. However, proteases are catalysts, and even low concentrations should eventually cleave and activate all of the receptors on a cell. How then do cells detect graded concentrations of a protease and respond in a concentration-dependent manner? Although low concentrations of thrombin will eventually cleave all PAR-1 at the cell surface, the rate of receptor cleavage correlates with the concentration of thrombin in the extracellular fluid (74). In particular, cumulative phosphatidylinositol hydrolysis correlates with cumulative receptor cleavage, suggesting that each cleaved and activated receptor generates a quantum of signal that is rapidly quenched. Thus cells detect different concentrations of peptides by graded receptor occupancy and detect different concentrations of proteases by the rate of receptor cleavage.

Evidence for Other PARs

Comparison of the structure and chromosomal location of PAR-1 and PAR-2 provides indirect evidence for the existence of a larger family of PARs. Although human PAR-1 and PAR-2 show only 35% identity at the protein level, their genes colocalize at chromosomal band 5q13, where the distance between these two genes is <100 kb (137, 138). The organization of human and rodent PAR-1 and PAR-2 genes is almost identical. The genes of both receptors contain two exons separated by an intron of 10-22 kb (Fig. 2). In both human genes, the first exons encode 29 residues, whereas the larger second exons contain the majority of the coding sequence and the protease cleavage site. This remarkably similar gene organization is indicative of a conserved evolutionary pattern from a common primordial gene and provides support for an extended gene family.

Differences in the potency with which the tachykinin family of neuropeptides stimulated various cell types and tissue preparations provided evidence for the existence of multiple receptors. These receptors were subsequently identified by molecular cloning (120). In a similar manner, comparisons of the biological activities of analogs of PAR-tethered ligands also suggests the existence of other receptors. For example, there are distinct differences in the potency with which analogs of the PAR-1-tethered ligand induce endothelium-dependent relaxation of rat aorta and stimulate contraction of rat gastric muscle, suggesting the existence of receptor subtypes (62). Similarly, differences in the potencies of PAR-1-derived peptides in human vascular tissues and in platelet aggregation assays suggest the presence of receptor subtypes (149). However, potency differences could also arise from differential degradation of peptides in different tissues or coupling to various signaling pathways, and proof of the existence of subtypes or new receptors requires molecular cloning.

    TRANSDUCING THE SIGNAL: MECHANISMS OF PAR SIGNAL TRANSDUCTION

Agonists interact with GPCRs, resulting in conformational changes that permit interaction of receptors with heterotrimeric G proteins, which catalyze the exchange of GDP by GTP on the alpha -subunit of the G protein. The alpha -subunit and the beta gamma heterodimer activate different effector enzymes or ion channels until GTP is hydrolyzed and the G proteins return to their inactive state (144). Among the three PARs, signal transduction pathways of PAR-1 have been the most extensively studied, and PAR-1 couples to a variety of G proteins, effector enzymes, and different signaling pathways, depending on the type of cell (55). This section discusses the mechanisms of signal transduction by PARs.

Signaling by PAR-1

Activation of heterotrimeric G proteins. In common with many GPCRs, PAR-1 couples to several different G proteins. The principal mechanism is through Galpha q proteins, resulting in activation of phospholipase C-beta , phosphoinositide hydrolysis, and formation of inositol trisphosphate and diacylglycerol, leading to Ca2+ mobilization and activation of protein kinase C (PKC; Fig. 5) (69). Activation of PAR-1 in CCL-39 fibroblasts, which naturally express the receptor, and in transfected cells stimulates formation of inositol trisphosphate in a manner that is mostly insensitive to pertussis toxin (69). Microinjection of antibodies that recognize Galpha q/11 into CCL-39 cells inhibits PAR-1-mediated Ca2+ mobilization and DNA synthesis (7), and Galpha q antibodies also inhibit PAR-1-stimulated GTPase activity in membranes prepared from platelets (11), confirming that PAR-1 couples to Galpha q/11. However, antibodies to Goalpha also inhibit PAR-1-mediated responses in CCL-39 cells (11). The inhibition suggests that PAR-1 also couples to Goalpha , which may account for the component of inositol trisphosphate formation that is sensitive to pertussis toxin (7, 69).


View larger version (24K):
[in this window]
[in a new window]
 
Fig. 5.   Mechanisms of signal transduction by PARs and receptor Tyr kinases. Most of this information derives from study of PAR-1. PLC, phospholipase C; PIP2, phosphatidylinositol 4,5,-bisphosphate; IP3, inositol trisphosphate; DAG, diacylglycerol; ERK, extracellular signal response kinase; MEK, mitogen-activated protein kinase kinase; PKC, protein kinase C; EGF, epidermal growth factor.

PAR-1 may also couple to other proteins of the Galpha q family, since thrombin stimulates incorporation of the photoreactive GTP analog [alpha -32P]GTP azidoanilide into Galpha 12 and Galpha 13 immunoprecipitated from platelet membranes (119). PAR-1 couples to Galpha 12 in 1321N1 astrocytoma cells to induce Ras-dependent activating protein 1 (AP-1)-mediated transcription and DNA replication, since thrombin-stimulated DNA synthesis is blocked by microinjection of antibodies to Galpha 12 (4). Similarly, Galpha 12 couples PAR-1 to AP-1-mediated gene expression when these proteins are coexpressed in COS-7 cells (126).

PAR-1 also couples to Gi proteins, which inhibit adenylyl cyclase and suppress formation of cAMP. Activation of PAR-1 in CCL-39 fibroblasts inhibits cAMP generation in a pertussis toxin-sensitive fashion, suggesting involvement of a Gi-like protein (69). Expression of a dominant negative mutant of Galpha i-2 in Chinese hamster ovary cells suppresses thrombin-stimulated arachidonic acid release, suggesting that Galpha i-2 couples PAR-1 to cytoplasmic phospholipase A2 (163).

The receptor domains that interact with heterotrimeric G proteins have been characterized by analysis of chimeras of PAR-1, the beta 2-adrenergic receptor (beta 2-AR), and the dopamine D2 receptor expressed in Xenopus oocytes and COS-7 cells (156). Replacement of the second intracellular loop of the Gs-coupled beta 2-AR or the Gi-coupled dopamine D2 receptor with the corresponding loop of the Gq-coupled PAR-1 is sufficient to switch the coupling of these receptors to Gq, resulting in inositol trisphosphate generation and Ca2+ mobilization. These results provide additional evidence that all the GPCRs undergo similar conformational changes to activate G proteins, regardless of the nature of their ligands.

In general, Galpha q subunits activate phospholipases beta 1 and beta 4, whereas the Gi beta gamma -subunits activate predominantly phospholipases beta 2 and beta 3 (44). Several observations indicate that PAR-1 activates phospholipase beta 1. Thrombin-stimulated inositol phosphate generation and Ca2+ mobilization are blunted in a mutant of the CCL-39 cell line that expresses low levels of phospholipase beta 1 but normal levels of the other phospholipases, suggesting that PAR-1 activates phospholipase beta 1 (45). In addition, activation of phospholipases A2 and D by thrombin is diminished in this cell line, suggesting that phospholipase beta 1 activates PKCalpha , which may be required for activation of phospholipases A2 and D (45). These findings are supported by the observations that overexpression of Galpha q, phospholipase beta 1, and phospholipase beta 2 in Xenopus oocytes enhances thrombin-induced Ca2+ release (28). Finally, thrombin stimulates phospholipase A2 activity and Na+/H+ exchange in platelets by activating PAR-1 (139).

Activation of Ras, Ras-related protein, and mitogen-activated protein kinase pathways. Recently, there has been considerable interest in the signaling mechanisms by which GPCRs, including PAR-1, stimulate DNA synthesis, cell growth, proliferation, and differentiation (14, 152). Point mutations of receptors that result in their constitutive activation in the absence of agonists are responsible for several genetically transmissible diseases. For example, mutations in the thyroid-stimulating hormone receptor result in thyroid adenomas, and constitutive activation of mutated luteinizing hormone receptor leads to agonist-independent transformation (122, 141).

Most of our understanding of how GPCRs regulate cell growth, division, and differentiation derives from the knowledge of viral oncogenes and their cellular homologues. These studies have focused on the role of receptor and nonreceptor Tyr kinases that regulate small G proteins, such as p21ras. The mechanisms by with epidermal growth factor (EGF) stimulates growth exemplifies this pathway (Fig. 5). EGF binding to its receptor results in receptor dimerization and autophosphorylation. The phosphotyrosine on the intracellular domain of the receptor binds through an SH2 domain to the adapter protein Shc, which recruits the Grb2-SOS complex to exchange GDP for GTP on p21ras. A phosphorylation cascade ensues: p21ras phosphorylates the Ser/Thr Raf-1 kinase [or mitogen-activated protein (MAP) kinase kinase kinase]. This kinase phosphorylates MEK-1/-2 (MAP kinase kinase), which in turn phosphorylates the extracellular signal response kinases (ERK)-1/-2 (MAP kinase). The end result is activation of transcription factors and consequent upregulation of immediate early genes that regulate growth.

The way GPCRs activate the MAP kinases pathway is complex and not completely understood (152). The situation is complicated by the fact that different GPCRs frequently couple several different G proteins, each of which may activate different pathways that merge with the receptor Tyr pathway at different levels. The Gi protein beta gamma complex activates nonreceptor Tyr kinases, such as Src, which might interact either with the Shc molecule or some receptor Tyr kinase (Fig. 5). Subsequent phosphorylation of Grb2-SOS complex activates the MAP kinases through p21ras. The role of Gs in regulation of MAP kinases pathway is unclear and might result from two antagonist pathways: similar to the Gi protein, the beta gamma -subunit of the Gs might activate the Shc-Grb2-SOS complex and p21ras, whereas cAMP release by Gs activation of the adenylyl cyclase may oppose this effect. Finally, Galpha q/11 also activates two different pathways generally defined as being dependent on or independent of Ras (Fig. 5). Activation of phospholipase C by Galpha q/11 results in phosphoinositide hydrolysis, Ca2+ mobilization, diacylglycerol formation, and PKC activation. PKC then phosphorylates and activates the Raf kinase. It has also been shown that activation of cellular PKC, calmodulin, and Ca2+ release also activates the Shc-Grb2-SOS complex, which stimulates MAP kinases through a Ras-dependent pathway.

Thrombin is mitogenic in many cells due to activation of PAR-1 and subsequent activation of MAP kinases (86, 100). Because PAR-1 couples to different G proteins (Galpha q/11, Go, Galpha 12, Galpha 13, and Gi), it is likely that PAR-1 activates MAP kinases through several pathways. In CCL-39 fibroblasts, which proliferate in response to thrombin, PAR-1 agonists activate the nonreceptor Tyr kinases Src and Fyn (31). Thrombin also induces Tyr phosphorylation of the adaptor protein Shc, which is then recruited to Grb2 (30). A dominant negative Shc that is deficient in Grb2 binding suppresses thrombin-stimulated activation of p44 MAP kinase and cell growth, indicating the importance of Shc in this pathway (30). In platelets, PAR-1 agonists activate p72syk and p60c-src, as determined by in vitro kinase assays, and stimulate their translocation to the cytoskeleton (134). In CCL-39 fibroblasts, thrombin activates p21ras in a manner that is inhibited by pertussis toxin and by the Tyr kinase inhibitor genistein, suggesting that activation of Ras involves Go and requires activation of protein kinases (153). Although the precise mechanism by which PAR-1 couples to Ras is still unclear, it is likely that Src and Fyn activate Ras through the adapter molecule Shc complexed with Grb2 and the SOS Ras exchange factor (30). It is still unknown whether this activation involves the beta gamma -subunit of Go. Besides these observations, it has been shown in platelets that PKC activates Ras, probably through a mechanism involving Galpha q/11 and a nonreceptor Tyr kinase (143).

Other small G proteins are involved in thrombin activation of cells expressing PAR-1. In platelets, the Ras-related protein Rap1B translocates to the cytoskeleton after activation by thrombin (49). Like thrombin-induced platelet aggregation, this translocation requires extracellular Ca2+. Rho is also involved in thrombin activation of neuronal cells (75). The mechanisms by which thrombin activates Rap1B and Rho remain to be elucidated. Moreover, differences in stimulation of CCL-39 fibroblasts by thrombin and the agonist peptide (159), and the finding that in platelets thrombin activates more p44 than p42 MAP kinase (121), suggest that p42 MAP kinase is more important in the regulation of cytoskeleton changes, whereas p44 MAP kinase is more involved in the mitogenic process. Finally, thrombin has been shown to stimulate c-fos (164) and the expression of the platelet-derived growth factor-A (78) in vascular smooth muscle cells.

In common with other GPCRs, thrombin also activates other enzymes that may contribute to the activation of MAP kinases. However, the function of some of these enzymes is not fully understood. In platelets, thrombin activates the phosphatidylinositide-3-kinase, which phosphorylates phosphatidylinositol 4,5-bisphosphate (79). Activation of Ras and the related protein Rho (57, 167), as well as the Src-related focal adhesion kinase are involved with phosphatidylinositide-3-kinase activation (27).

PAR-2 and PAR-3 Signal Transduction

Compared with PAR-1, very little is known about mechanisms of signal transduction that couple to PAR-2 and PAR-3. Agonists of PAR-2 stimulate generation of inositol trisphosphate and mobilization of intracellular Ca2+ in numerous cell types, including transfected cell lines, epithelial cells (enterocytes and keratinocytes), smooth muscle cells, and various tumor cell lines (18, 35, 83, 115-117, 133). Therefore, it is likely that PAR-2 couples to Galpha q or Galpha o. In enterocytes and transfected KNRK kidney epithelial cells, agonists of PAR-2 also stimulate arachidonic acid release and rapid generation of prostaglandins E2 and F1alpha , suggesting activation of phospholipase A2 and cyclooxygenase-1 (83). In smooth muscle cells and enterocytes, PAR-2 agonists also strongly activate MAP kinases ERK-1/-2 and weakly stimulate the MAP kinase homologue p38, although c-Jun NH2-terminal kinase is not activated (10, 165, and Bunnett, unpublished observations). Trypsin and agonist peptides also induce c-Fos and cyclooxygenase-2 in enterocytes (Bunnett, unpublished observations). PAR-2 activates the protein-Tyr phosphatase SHP2 involved in the regulation of Src phosphorylation (165).

    PHYSIOLOGICAL AND PATHOPHYSIOLOGICAL FUNCTIONS OF PAR

The requirements for PAR activation are the presence of an active enzyme in the extracellular fluid at sufficiently high concentrations and the existence of uncleaved PAR at the surface of target cells where it can be activated by proteolysis and couple to signaling pathways. The presence of active proteases depends on release or generation and on the presence of protease inhibitors, and the availability of PAR at the cell surface is governed by trafficking of the receptor from intracellular stores and on the presence of G proteins and G protein receptor kinases that modify activity.

There is little doubt that thrombin is the physiologically relevant activator of PAR-1 and PAR-3. Thrombin is widely distributed because it is a component of the coagulation cascade, and PAR-1 and PAR-3 are present on multiple cell types where they could be activated by thrombin. Pancreatic trypsin is the most efficient activator of PAR-2, but there is a discrepancy between the availability of pancreatic trypsin and the distribution of PAR-2. Biologically active trypsin is present in the lumen of the small intestine, where it may activate PAR-2 at the apical membrane of enterocytes (83), but PAR-2 is also found in many tissues where it must be activated by other proteases with trypsinlike specificity. These remain to be identified.

There are formidable problems in defining the physiological role of PARs. The biological roles of GPCRs for neurotransmitters and hormones are usually defined by examining the effects of potent and selective receptor agonists or antagonists and, more recently, by genetic deletion of the agonist or its receptors. At present, many of these tools are not available for PARs. The most efficient agonists are proteases, but these cleave proteins other than their receptors, and this cleavage may also mediate biological effects. Peptides corresponding to the tethered ligands may be more selective agonists. However, some of these peptides activate more than one PAR, and they activate receptors with very low potency compared with proteases (13, 135). Antagonists of PAR-1 are available (12), but they do not presently exist for PAR-2 and PAR-3. Several groups have generated mice in which the genes encoding PAR-1 and PAR-2 are deleted by homologous recombination (34, 38), although, unless there is an obvious abnormal phenotype, it is challenging to identify altered functions. Despite these difficulties, considerable progress has been made in defining the physiological and pathophysiological roles of PARs.

Functions of PAR-1

Thrombin is generated after vascular injury as part of the coagulation cascade. Thrombin cleaves fibrinogen to form fibrin, which is the basis of the blood clot and which also activates the clotting factors V, VIII, XIII, and protein C. However, in addition to a central role in coagulation, thrombin has numerous biological functions that are related to inflammation, tissue remodeling, and healing. Many of these effects are mediated by PAR-1.

PAR-1 distribution. The distribution of PAR-1 in tissues and cell lines has been evaluated by several techniques, including Northern blotting, RT-PCR, in situ hybridization, and immunohistochemistry. PAR-1 mRNA is expressed in platelets, endothelial cells, fibroblasts, monocytes, T cell lines, osteoblast-like cells, smooth muscle cells, certain epithelial cells, uterine stromal cells, neurons and glial cells in the brain, and certain tumor cell lines (5, 33, 56, 65, 76, 112, 160, 162). Analysis of rat brain by in situ hybridization and immunohistochemistry revealed high expression of PAR-1 in neurons of the neocortex, cingulate cortex, subsets of thalamic and hypothalamic nuclei, and discrete layers of the hippocampus, cerebellum, and olfactory bulb, as well as by astroglia (111, 162).

Expression of PAR-1 is altered in certain diseases. Thus analysis by in situ hybridization and immunohistochemistry indicated that, whereas PAR-1 is mostly confined to the endothelium of normal human arteries, in atheroma it is highly expressed in regions rich in macrophages, smooth muscle cells, and mesenchymal-like intimal cells (108). Therefore, PAR-1 may contribute to sclerotic and inflammatory processes in the human vasculature. PAR-1-expressing cells are also abundant in the synovial membrane of patients with rheumatoid and osteoarthritis, where PAR-1 may contribute to the inflammatory process (101, 102, 142).

PAR-1 and inflammation. Many of the biological actions of PAR-1 are related to inflammation. In addition to the critical role of PAR-1 and thrombin in platelet aggregation (68), proinflammatory effects of thrombin that are mediated by PAR-1 include 1) vasodilatation and vasoconstriction, 2) increased vascular permeability, and 3) cellular adhesion and infiltration by chemotaxis.

The effects of thrombin on the vasculature have been examined in many species and in preparations ranging from isolated tissues in organ baths to the intact animal. Thrombin may regulate blood flow by direct effects on the vasculature or by the release of vasoactive substances from other cells. The direct effects have been examined in isolated tissues. The consensus of these studies is that thrombin and peptides corresponding to the tethered ligand of PAR-1 affect vascular tone in two ways: 1) they relax precontracted tissues by an endothelial-dependent mechanism that involves release of nitric oxide (104), and 2) they contract vascular smooth muscle by a direct effect that requires extracellular Ca2+ (3, 39, 104). Thus thrombin and peptides corresponding to the PAR-1-tethered ligand induce relaxation of isolated coronary arteries from dog (85) and pig (150), relax rat aorta (104), and contract human placental and umbilical arteries (149). Infusion of the PAR-1 agonist peptide SFLLRN into the coronary artery of an anesthetized dog causes a transient increase followed by a sustained decrease in blood flow in this vessel (37). In addition to the direct effects of thrombin on the vasculature, thrombin stimulates release of serotonin from platelets (6) and histamine from mast cells (127), which both cause vasodilatation. Thus thrombin generation at a site of vascular injury may affect vascular tone and blood flow by activating PAR-1 on multiple cell types.

Increased vascular permeability to plasma proteins and resultant edema and influx of inflammatory cells are critical early steps of inflammation. Thrombin and PAR-1 agonists affect vascular permeability by direct effects on the vasculature and by releasing mediators from other cells. The direct effects have been studied using isolated endothelial cells. Thrombin causes rapid and transient contraction of endothelial cells from the human umbilical vein, resulting in gap formation and increased permeability of confluent cells (87). Exposure of the subendothelium may be conducive to atherogenesis and thrombosis. Similarly, thrombin increases transendothelial clearance of albumin in confluent monolayers of endothelial cells from bovine pulmonary artery with a half-life of ~1 min, which coincides with reorganization of actin filaments (91). Increased permeability to albumin is immediately preceded by generation of inositol trisphosphate and mobilization of Ca2+ and is probably mediated by PAR-1. The role of thrombin and PAR-1 in edema has also been examined in intact animals (32). Injection of thrombin and PAR-1 agonist peptides into rat paw induces edema and extravasation of plasma proteins, mediated in part by release of biogenic amines released from mast cells, including serotonin and histamine. Furthermore, the thrombin inhibitor hirudin attenuates carrageenin-induced edema of the rat paw, suggesting that thrombin acts as an inflammatory mediator in vivo by activating PAR-1 on mast cells to stimulate release of vasoactive amines. Indeed, thrombin stimulates mast cell degranulation, releasing inflammatory mediators such as histamine (127), and also enhances production of interleukin-1 by macrophages treated with lipopolysaccaride (77). Interleukin-1 has multiple biological actions and plays a major role in host defense in inflammation, injury, and trauma.

Extravasation of leukocytes is a critical step in the genesis of inflammation. Thrombin induces expression of chemotactic agents by activating PAR-1. Monocyte chemotactic protein-1 is a potent and specific chemotactic factor for monocytes that is involved in cellular immune reactions and responses to acute tissue injury. Thrombin activates PAR-1 on endothelial cells to induce expression and release of monocyte chemotactic protein-1, which stimulates monocyte chemotaxis (33). Thrombin also promotes neutrophil adhesion transendothelial passage (8, 103) and stimulates expression of P-selectin by endothelial cells (146). Together, these findings indicate that PAR-1 is important in recruitment of monocytes during vascular injury.

PAR-1 and mitogenesis. As discussed in Activation of Ras, Ras-related protein, and mitogen-activated protein kinase pathways, PAR-1 agonists activate signaling pathways that are important for growth and differentiation, and thus it is not surprising that PAR-1 agonists are mitogenic in many cells. Thrombin and peptides corresponding to the tethered ligand of PAR-1 stimulate proliferation of vascular smooth muscle cells (78, 93, 100, 164), fibroblasts (31, 153), endothelial cells (96, 136), and mesangial cells (2). These effects of thrombin are important, for they permit healing and repair of inflamed and damaged tissues.

Observations in mice in which the gene encoding PAR-1 is deleted by homologous recombination suggest that this receptor also plays a role in embryogenesis (34, 38). Thus deletion of PAR-1 results in 50% mortality at embryonic days 9-10, whereas one-half the animals proceed to term and become grossly normal. Despite the widespread distribution of PAR-1 in the cardiovascular system, these knockout mice had surprisingly normal cardiovascular parameters. Thus PAR-1 plays a critical, but as yet undefined, role in development.

PAR-1 is also expressed in several tumor cell lines, in which PAR-1 activation enhances adhesion to platelets, fibronectin, von Willebrand factor, and endothelial cells (81, 112, 113). Thus thrombin and PAR-1 may contribute to tumor adhesion and subsequent metastasis. Consequently, PAR-1 antagonists may be useful in preventing or minimizing the metastatic events.

PAR-1 in the nervous system. Recent observations suggest an important role for thrombin and PAR-1 in the brain under normal conditions and following injury (48, 55).

Although most prothrombin is produced by the liver, prothrombin mRNA is present in the normal brain, including the olfactory bulb, the cortex, and the cerebellum (41). PAR-1 is also widely expressed in the central nervous system. PAR-1 mRNA is highly expressed in the brain and dorsal root ganglia of the neonatal rat, although levels decline after birth (111). At postnatal day 28, PAR-1 is expressed in the substantia nigra, the ventral tegmental area, the pretectal area, certain hypothalamic nuclei, and certain regions of the cortex. It is localized to neurons, glial cells, and ependymal cells, but the white matter and most endothelial cells do not express detectable PAR-1 (162). A comparison of the distribution of prothrombin and PAR-1 in the brain indicates that there is a distinct, but in some regions overlapping, localization. In many instances, prothrombin and PAR-1 are expressed in similar regions, suggesting that thrombin may regulate cells expressing PAR-1 in a paracrine or even autocrine manner under normal circumstances. For example, in the cerebellum PAR-1 is expressed in the Purkinje cell layer, and prothrombin is found in the Purkinje and granule cell layers. In addition, thrombin would be expected to enter the brain from the circulation when the blood-brain barrier is disrupted by trauma and could thus have widespread effects on many cell types expressing PAR-1.

One of the most striking consequences of PAR-1 activation is to induce morphological alterations in neurons and glial cells. Thrombin causes retraction of processes by neuronal cells (58). Similarly, thrombin and PAR-1 agonist peptides cause astrocytes to lose their stellate morphology and become flat and epithelial in shape (9, 26). Other effects of thrombin on astrocytes include release of the vasoconstrictor endothelin-1 (42), increased synthesis and secretion of nerve growth factor (109), and reduced expression of the metabotropic glutamate receptor (95). Thrombin is also mitogenic for astrocytes (123, 124). In view of the important role of astrocytes in support of neurons and the blood-brain barrier, these effects of thrombin may have widespread implications. Notably, thrombin and SFLLRN protect both astrocytes and neurons in culture from cell death induced by environmental insults such as hypoglycemia and oxidative stress (154). Because thrombin would likely enter the brain on damage, it may play an important role in protecting the astrocytes and neurons from death. Furthermore, PAR-1 agonists have been reported to enhance and attenuate the neurotoxicity of beta -amyloid, a protein implicated in Alzheimer's disease (125, 145). This observation suggests a role for thrombin and PAR-2 in the pathogenesis of neural injury.

The biological effects of thrombin are modulated by the protease nexin-1, a Ser protease inhibitor that has high affinity for thrombin and that is found mostly in the brain (48). Protease nexin-1 is highly expressed in the brain and is localized to areas that also express prothrombin and would be exposed to thrombin. Therefore, the biological actions of thrombin in the brain are inhibited by protease nexin-1, which counteracts the effects of thrombin on morphology, growth, and protection from environmental insults (9, 26). Protease nexin-1 expression is diminished around blood vessels in Alzheimer's disease (155).

Functions of PAR-2

In comparison with PAR-1, we know very little about the physiological and pathophysiological importance of PAR-2. One of the major obstacles to defining potential functions is that, in most tissues, the enzymes that cleave and activate this receptor remain to be identified.

PAR-2 distribution. PAR-2 is expressed in the gastrointestinal tract, pancreas, kidney, liver, airway, prostate, ovary, and eye (18, 115-117) and is found in epithelial and endothelial cell lines, smooth muscle, T cell lines, neutrophils, and certain tumor cell lines (1, 18, 35, 66, 70, 83, 92, 96, 132, 133).

Although nanomolar concentrations of trypsin cleave and activate PAR-2, the widespread distribution of PAR-2 compared with the relatively limited distribution of pancreatic trypsin suggests that other trypsinlike enzymes activate PAR-2 in some locations. Pancreatic trypsin may activate PAR-2 in some tissues under physiological and pathophysiological conditions. Thus trypsin in the lumen of the small intestine may activate PAR-2 on enterocytes (18, 83), and trypsin may activate PAR-2 in the pancreas during pancreatitis, when there is premature activation of trypsinogen (18 and T. Nguyen, M. Moody, C. Okolo, and N. Bunnett, unpublished observations). Elsewhere, other trypsinlike enzymes probably cleave and activate PAR-2. Trypsinogen-2 mRNA and its protein product are expressed by endothelial cells (84), and some types of human cancer cells secrete trypsins (84) that may activate PAR-2. Another candidate is mast cell tryptase, a major secretory granule protein of mast cells. Tryptase cleaves and activates PAR-2 in transfected cell lines and in cells that naturally express this receptor (35, 98).

PAR-2 and inflammation. PAR-2, like PAR-1, has many effects that are proinflammatory. Thus trypsin and peptides corresponding to the tethered ligand of PAR-2 induce relaxation of rings of rat aorta, which is dependent on nitric oxide production from the endothelium (1, 132). However, in the absence of the endothelium, PAR-2 agonists cause neither contraction nor relaxation (1, 132). Similarly, trypsin and PAR-2 peptides elicit endothelium-dependent, nitric oxide-mediated relaxation of the porcine coronary artery but do not cause contraction in the absence of the endothelium (70). In contrast, trypsin stimulates contraction of rabbit aorta in the absence of the endothelium, and induces Ca2+ mobilization in isolated myocytes (82). It is possible that myocytes from rabbit aorta express PAR-2, in contrast to those from rat aorta, or that trypsin activates another receptor in the rabbit. In the intact rat, intravenous injection of SLIGRL produces a marked fall in blood pressure, consistent with release of nitric oxide from endothelial cells (70).

The involvement of PAR-2 in inflammation is supported by the finding that PAR-2 mRNA is unregulated by tumor necrosis factor-alpha and interleukin-1alpha , both of which act together to orchestrate the acute inflammatory response by either inducing or reducing the transcription of a large number of genes in responding cells (118).

Mast cell tryptase cleaves and activates PAR-2 in many cell types, including transfected cells, endothelial cells, enterocytes, and colonic myocytes (35, 98). The observation that tryptase cleaves and activates PAR-2 suggests a role for this receptor in inflammatory states characterized by mast cell infiltration and degranulation. Tryptase is released by mast cells when they degranulate in response to inflammatory stimuli and has many effects that appear to be receptor mediated. Tryptase is mitogenic for epithelial cells, fibroblasts, and smooth muscle cells and stimulates intracellular adhesion molecule expression by epithelial cells (23-25, 59, 131), but the receptor that mediates these effects has not been identified. Although PAR-2 is expressed by several cell types that respond to tryptase, including epithelial, endothelial, and smooth muscle cells, proof that PAR-2 mediates the effects of tryptase requires the use of selective agonists and antagonists.

Mitogenesis. Like PAR-1, PAR-2 agonists stimulate MAP kinase activity and may regulate cellular growth and proliferation (10, 165). However, the effects of PAR-2 agonists on growth depend on the cell type. Both PAR-2 and PAR-1 agonists stimulate proliferation of endothelial cells (96). In contrast, whereas PAR-1 agonists stimulate growth of keratinocytes, PAR-2 agonists inhibit keratinocyte growth and differentiation (40). In addition, PAR-2 agonists inhibit colony formation by tumor cell lines (18).

PAR-2 in the gastrointestinal tract. PAR-2 is highly expressed by enterocytes, where it is localized to the apical and basolateral membranes (Fig. 6, A and B). Trypsinogen is secreted into the intestinal lumen after feeding, where it is activated by enterokinase. Although trypsin is traditionally regarded as an enzyme with the principal function of degrading dietary proteins, recent observations indicate that it may also regulate enterocytes by cleaving and triggering PAR-2 at the apical membrane (83). Thus physiological concentrations of trypsin as well as PAR-2 agonist peptides stimulate generation of inositol trisphosphate and mobilization of intracellular Ca2+, induce arachidonic acid release, and cause generation of prostaglandins E2 and F1alpha (Fig. 6C). These effects occur when agonists are applied selectively to either the apical or basolateral membranes of enterocytes, proving functional evidence for the expression of PAR-2 at both surfaces. Although the physiological consequences of this stimulation remain to be determined, eicosanoids regulate multiple processes within the intestine, suggesting that trypsin may influence intestinal function through PAR-2 activation and prostaglandin release.


View larger version (92K):
[in this window]
[in a new window]
 
Fig. 6.   A and B: localization of PAR-2 to enterocytes of the rat small intestine by indirect immunofluorescence and confocal microscopy. A: cross section through the base of the crypts (*, lumen). B: section through a villus tip. PAR-2 immunoreactivity at the apical membrane is indicated by arrowheads and at the basolateral membrane by arrows. C: effects of PAR-2 agonists on Ca2+ mobilization in enterocytes. Data are from Refs. 18 and 83.

However, PAR-2 is expressed by cell types and in locations in the gastrointestinal tract that are not exposed to luminal trypsin. PAR-2 is highly expressed by gastrointestinal smooth muscle and at the basolateral membrane of enterocytes (Fig. 7) (35, 83). The enzymes that activate PAR-2 at these locations are unknown, but one possibility is mast cell tryptase. Indeed, tryptase activates PAR-2 on enterocytes and colonic myocytes (Fig. 7). PAR-2 activation inhibits the amplitude of rhythmic contractions of strips of rat colon, which is unaffected by indomethacin, L-NG-nitroarginine methyl ester, a B2 receptor antagonist, and tetrodotoxin (35). PAR-2-mediated inhibition of motility may contribute to motility disturbances of the colon during conditions associated with mast cell degranulation.


View larger version (56K):
[in this window]
[in a new window]
 
Fig. 7.   A: localization of PAR-2 to myocytes of rat colon by indirect immunofluorescence and confocal microscopy. PAR-2 immunoreactivity in the muscularis mucosa is indicated by arrowheads and in the circular muscle by arrows. muc, Mucosa; sm, submucosa; cm, circular muscle. B: effects of PAR-2 agonists on Ca2+ mobilization in cultured myocytes. Data are from Ref. 35.

The finding that PAR-2 is highly expressed in the pancreas is intriguing, since this is the principal site of trypsin production. However, under normal circumstances very little active trypsin is secreted by acinar cells, the principal product being the inactive zymogen trypsinogen. Trypsin is prematurely activated in the inflamed pancreas, where it may activate PAR-2. PAR-2 is expressed by both major exocrine cell types of the pancreas: acinar cells that release digestive enzymes and duct cells that produce fluid and bicarbonate. Trypsin and PAR-2 agonist peptides induce Ca2+ mobilization and stimulate amylase release from isolated pancreatic acini (18 and Bunnett, unpublished observations). Application of these agonists to the basolateral but not apical membrane of monolayers of pancreatic duct cells increases short-circuit current, due to activation of Ca2+-sensitive Cl- and K+ channels (T. Nguyen, M. Moody, C. Okolo, and N. Bunnett, unpublished observations). These effects may be of relevance in pancreatitis, when trypsin is released across the basolateral membrane. Trypsin is also released into the circulation during pancreatitis, raising the possibility that PAR-2 may mediate some of the systemic actions of trypsin in this disease.

Functions of PAR-3

PAR-3 is the second thrombin receptor (71). In humans it 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, PAR-3 is expressed by megakaryocytes. Although PAR-3 may mediate the effects of thrombin in some sites, its biological role is presently unknown.

    TURNING OFF THE SIGNAL: MECHANISMS OF PAR INACTIVATION

Cellular responses to Ser proteases and to most agonists of GPCRs are rapidly attenuated (16). The mechanisms that attenuate signaling by Ser proteases are of interest for several reasons. First, attenuation of signal transduction is important to prevent the uncontrolled stimulation of cells, which may otherwise result in dysfunction and possibly disease. Second, in view of the unusual mechanism of receptor activation, defining the mechanisms of attenuation of PAR signaling provides novel insight into the regulation of GPCRs in general. Finally, an understanding of mechanisms of attenuation may be invaluable in designing long-acting antagonists and agonists of PARs that may be therapeutically useful.

The mechanisms of signal attenuation have been most thoroughly investigated for the GPCRs for hormones and neurotransmitters (16). In common with many important biological processes, there are multiple parallel processes that attenuate signaling by these receptors and considerable apparent redundancy (Fig. 8). Thus signals are attenuated by mechanisms that operate at the level of the agonist, the receptor, and the G proteins and at numerous downstream steps of the signaling pathway. Agonist removal from the extracellular fluid is the earliest mechanism of attenuation. Several processes remove hormones and neurotransmitters from the extracellular fluid, including diffusion and dilution in interstitial fluid, uptake by high-affinity transporters, and enzymatic degradation. Receptor desensitization occurs during short-term (seconds to minutes) exposure of cells to agonists and is mediated by uncoupling of activated receptors from G proteins to terminate signaling. Receptor endocytosis depletes the plasma membrane of high-affinity receptors and contributes to both desensitization and resensitization of signaling. Receptor downregulation is a loss of receptors from a cell that results from long-term (hours to days) continuous exposure of cells to agonists.


View larger version (21K):
[in this window]
[in a new window]
 
Fig. 8.   Potential mechanisms of agonist-induced desensitization and trafficking of PARs. Most of the information used to make this schematic derives from studies of PAR-1, although desensitization and trafficking of PAR-2 has been also investigated, but in less detail. 1) PARs are activated by irreversible proteolytic cleavage, which exposes the tethered ligand domain and permits it to interact with the cleaved receptor. However, the cleaved receptor cannot be reactivated by proteolysis. 2) PAR activation presumably induces targeting of G protein receptor kinases (GRKs) and beta -arrestins to the receptor in the plasma membrane, as for other GPCRs. 3) Phosphorylation by GRKs and second messenger kinases and interaction with beta -arrestins may uncouple the receptor from G proteins and quench the signal. 4) The receptor is internalized by a clathrin-mediated mechanism. beta -Arrestins may serve as adaptor molecules for clathrin. 5) The internalized PAR is targeted to lysosomes. Some PAR recycles, but would not be active. 6) There are large intracellular pools of PAR in the Golgi apparatus and in vesicles. 7) Resensitization involves mobilization of these pools.

Less is known about the mechanisms that attenuate signal transduction by PARs. However, there are informative differences and similarities in these processes for PARs and for GPCRs for peptide and nonpeptide hormones and neurotransmitters. In this section we review what is known about mechanisms of attenuation of PAR signaling and highlight differences and similarities in these mechanisms between PARs and other GPCRs.

PAR Cleavage

Soluble agonists of GPCRs are removed from the extracellular fluid by diffusion, uptake, and enzymatic degradation. Together, these processes comprise one of the earliest mechanisms of signal attenuation. Similarly, proteases are removed from the vicinity of their receptors by diffusion and may be inactivated by protease inhibitors and degradation. However, once cleaved and activated, the tethered ligand is always available to bind to the receptor and cannot be removed by diffusion and dilution or by uptake. The continued presence of an agonist that is physically part of the receptor presents unique problems for terminating the signal.

Although proteolysis is the physiological mechanism for PAR activation, proteases can also inactivate receptors. Cleavage of PARs could remove the tethered ligand before it is exposed, and thus generate a receptor that is unresponsive to an activating protease, or could inactivate the tethered ligand after exposure, and thereby terminate the signal. As discussed in Deactivation by Other Proteases, cathepsin G, elastase, and proteinase 3 cleave PAR-1 to remove the tethered ligand domain and thereby render a receptor that is unresponsive to thrombin (99, 128). Cleavage of the tethered ligand after it is exposed may also terminate the signal, since thermolysin attenuates the otherwise sustained Ca2+ response of SF9 insect cells expressing PAR-1 to thermolysin (29).

The irreversible nature of PAR activation by proteolytic cleavage renders cleaved receptors resistant to further proteolytic activation. Thus cleavage of surface receptors by thrombin and trypsin renders cells unresponsive to proteases for considerable periods, until the plasma membrane is replenished with intact receptors.

Proteolysis is also important for terminating signaling by neuropeptides. In a manner that is reminiscent of the role of acetylcholinesterase in terminating cholinergic transmission in the neuromuscular junction of skeletal muscle, cell surface proteases degrade and inactivate neuropeptides in the extracellular fluid. Neutral endopeptidase (EC 3.4.24.11) is a prototypical neuropeptide-degrading enzyme that resembles thermolysin (129). It is anchored to cells by a single transmembrane domain, and the bulk of the protein, including the active site, projects into the extracellular fluid, where it is well placed to degrade neuropeptides, notably substance P, and thus terminate signaling.

PAR Desensitization

The early events of signaling by GPCRs, exemplified by generation of second messengers such as cAMP and inositol 1,4,5-trisphosphate, are usually rapidly attenuated by regulatory processes collectively known as receptor desensitization. Two types of desensitization can be distinguished on the basis of the underlying mechanism. Homologous desensitization is mediated by agonist-dependent activation of the same receptor, whereas heterologous desensitization is caused by activation of a different receptor. An important component of desensitization is uncoupling of the activated receptor from its G proteins by receptor phosphorylation (Fig. 8) (16). Two classes of protein kinases mediate this phosphorylation. A unique class of Ser/Thr protein kinases, the G protein receptor kinases (GRKs), mediate agonist-dependent phosphorylation of GPCRs and initiate homologous desensitization. This desensitization also depends on the arrestins, which serve as functional cofactors for GRKs. Second messenger-dependent kinases [protein kinases C (PKC) and A (PKA)] mediate agonist-independent phosphorylation of receptors and initiate heterologous desensitization.

GRK-induced desensitization. GRKs and beta -arrestins are of established importance for desensitization of many GPCRs for hormones and neurotransmitters, most notably the beta 2-AR (16). In the presence of agonists, GRK-2 and GRK-3 phosphorylate the beta 2-AR. beta -Arrestins serve as functional cofactors for the GRKs by binding to phosphorylated receptors and disrupting the interaction between the receptor and G proteins.

Less is known about the role of GRKs and beta -arrestins in desensitization of PARs. However, uncoupling of the activated PARs from G proteins is likely to be of particular importance, because, once cleaved, the tethered ligand is always exposed, which would result in continuous or prolonged stimulation unless there are efficient mechanisms for terminating the signal. Indeed, cellular responses to activation of PAR-1 and PAR-2 are transient and undergo homologous desensitization to repeated stimulation. This observation indicates the existence of efficient mechanisms that turn off signaling.

PAR-1 and PAR-2 have several potential consensus sites for GRK phosphorylation in the COOH-tail and third transmembrane domains. Thrombin causes rapid phosphorylation of PAR-1 in transfected cells (72, 157, 158). Although activation of PKC with phorbol esters also induces phosphorylation of PAR-1, thrombin-stimulated phosphorylation persists after inhibition or downregulation of PKC, suggesting that PKC is not principally responsible for agonist-induced PAR-1 phosphorylation (72). Rather, GRKs are mostly responsible for agonist-induced PAR-1 phosphorylation, since thrombin-stimulated phosphorylation is elevated by coexpression of GRK-3 (72). Indeed, coexpression of GRK-3 with PAR-1 in Xenopus oocytes inhibits responses to EC50 concentrations of thrombin, whereas GRK-2 is 10- to 25-fold less effective. Although the site of phosphorylation has not been precisely mapped, domains in the COOH-tail are important, since truncation of the COOH-tail at Cys387, and substitution of all the Ser and Thr residues in the COOH-tail to Ala render PAR-1 insensitive to regulation by GRK-3 (72). Furthermore, thrombin signals more robustly in cells expressing this Ser/Thrright-arrowAla mutant or PAR-1 truncated at Tyr397 compared with wild-type PAR-1, which indicates diminished desensitization (140).

The importance of the COOH-tail in desensitization is illustrated by comparison of agonist-induced phosphorylation and desensitization of inositol phosphate generation of cells expressing PAR-1, the 5-hydroxytryptamine (5-HT2) receptor, or a chimeric receptor (157). Thrombin induces rapid phosphorylation and extensive desensitization of PAR-1, whereas serotonin does not induce phosphorylation of the 5-HT2 receptor and causes a far slower and smaller desensitization of its receptor. However, a chimeric 5-HT2 receptor carrying the PAR-1 COOH-tail undergoes serotonin-stimulated phosphorylation and desensitization. Therefore, the COOH-tail of PAR-1 contains motifs that trigger uncoupling from G proteins and consequent receptor desensitization.

It is probable that beta -arrestins interact with GRK-phosphorylated PARs and thereby contribute to desensitization. However, as yet the role of beta -arrestins in PAR-1 desensitization and the involvement of GRKs and beta -arrestins in PAR-2 and PAR-3 desensitization are unknown.

Second messenger kinase-induced desensitization. Second messenger kinases also contribute to desensitization of receptors for hormones and neurotransmitters such as the beta 2-AR (16). There are multiple PKC consensus sites in the COOH-tail and third intracellular loop of PAR-1 and PAR-2, suggesting that PKC may phosphorylate these receptors and contribute to desensitization. In support of this possibility, phorbol esters stimulate phosphorylation and inhibit signaling of PAR-1 (72). The observation that signaling of GRK-3-insensitive mutants of PAR-1 is still rapidly attenuated suggests that PAR-1 also relies on an additional desensitization mechanism, perhaps using PKC sites in the intracellular loops (72). Acute activation of PKC also inhibits PAR-2 activation in transfected epithelial cells and enterocytes, whereas PKC inhibition magnifies these responses (17). Together, these observations suggest that PKC regulates both PAR-1 and PAR-2, but it is not known whether PKC acts at the level of the receptor or at steps that are downstream in the signaling cascade. The existence of several parallel mechanisms for PAR desensitization involving both GRKs and second messenger kinases is not unreasonable, since both GRKs and PKA contribute to desensitization of the beta 2-AR. The relative contribution of these kinases depends on the type of desensitization (homologous or heterologous) and on the concentration of agonists (GRKs mediate desensitization to high and PKA to low agonist concentrations).

PAR Endocytosis and Trafficking

The ability of cells to respond to proteases, hormones, and neurotransmitters requires the presence of GPCRs that are appropriately located at the plasma membrane where they can interact with high affinity with hydrophilic agonists in the extracellular fluid. Thus the responsiveness of target cells is critically dependent on the subcellular distribution of receptors, and agonist-induced endocytosis and trafficking of GPCRs is important in desensitization and resensitization of signaling. Although endocytosis of cell surface receptors and subsequent intracellular sorting are critically important cellular processes, the requirements for endocytosis, the endocytic mechanism, the fate of the internalized receptor and ligand, and the function of internalization and subsequent trafficking differ between receptors.

Pathways of endocytosis and trafficking. Most information about the pathways of agonist-induced trafficking of GPCRs derives from studies of receptors for hormones and neurotransmitters (16). For example, substance P induces clathrin-mediated endocytosis of the NK1-R in transfected epithelial cells, in primary cultures of enteric neurons, and in endothelial cells and neurons of the intact animal. Substance P and the NK1-R internalize into the same vesicles, which also contain the transferrin receptor and are thus early endosomes. Endosomal acidification permits dissociation of the ligand and its receptor and sorting into different pathways: substance P is degraded in lysosomes, and the NK1-R recycles to the cell surface, where it may respond again to substance P. Agonists have similar effects on the subcellular distribution of the beta 2-AR and the receptor for gastrin-releasing peptide.

There are similarities and distinct differences in endocytosis and trafficking of PARs and receptors for neurotransmitters and hormones. After activation by proteolysis or by activating peptides, PAR-1 and PAR-2 rapidly internalize (17, 22, 60, 67). In erythroleukemic or megakaryoblastic cell lines, ~85% of immunoreactive PAR-1 is sequestered from the cell surface into coated pits and then into endosomes within 1 min of activation. The initial accumulation of PAR-1 into coated pits suggests that internalization proceeds by a clathrin-mediated process (67). Similarly, trypsin and activating peptides trigger rapid endocytosis of PAR-2 in transfected epithelial cells (17). PAR-1 and PAR-2 initially internalize into early endosomes that contain the transferrin receptor. However, in contrast to the NK1-R and beta 2-AR, which mostly remain in endosomes, the bulk of internalized PAR-1 and PAR-2 quickly traverse the endosomal network and are sorted to lysosomes (Fig. 8). Within 30-60 min of stimulation with proteases or activating peptides, PAR-1 and PAR-2 are detected in lysosomes in transfected cells and cell lines that naturally express these receptors. However, the fidelity of lysosomal targeting is not complete, because a portion of activated and internalized PARs also recycle to the plasma membrane (22, 67). As expected, these cleaved and recycled receptors are unable to respond to thrombin, although they can be activated by peptide agonists. Therefore, the physiological relevance of this recycling is unknown.

Recent evidence suggests that PAR-1 internalizes constitutively, even in the absence of ligand, and shuttles back and forth between the plasma membrane and the intracellular stores (140). When surface receptors were labeled with an antibody to an extracellular domain at 4°C and then warmed to 37°C, there was a loss of surface PAR-1 that coincided with the appearance of labeled receptors in the intracellular pool; reappearance of tagged receptors at the cell surface occurred when surface antibodies were removed. Together, these findings indicate that PAR-1 tonically travels between the plasma membrane and intracellular stores. Some single transmembrane domain receptors, notably the transferrin receptor, also constitutively internalize and recycle in the absence of ligand. However, the beta 2-AR or NK1-R internalize and recycle only in the presence of agonist.

In contrast to the beta 2-AR and the NK1-R, which are mostly confined to the plasma membrane of transfected cells and cells that naturally express these receptors in the absence of agonists, there are extensive intracellular pools of PAR-1 and PAR-2 in certain cells. Endothelial cells and fibroblasts have large intracellular pools in the Golgi apparatus and in a distinct tubulovesicular network that contains the transferrin receptor and resembles the tubulovesicular endosomal reticulum (60, 64). Platelets also have an intracellular store of PAR-1 (97). In transfected cells and enterocytes, there are prominent pools of PAR-2 in the Golgi apparatus (17, 83).

Function of agonist-induced endocytosis and trafficking. Endocytosis depletes the plasma membrane of high-affinity receptors that are capable of interacting with hydrophilic agonists in the extracellular fluid and may thereby contribute to receptor desensitization (16). However, this is not the principal mechanism, since responses to agonists of the NK1-R and beta 2-AR still desensitize even after inhibition of endocytosis of these receptors by pharmacological means or by receptor mutation. Instead, endocytosis, receptor processing, which may include dissociation of ligand and beta -arrestins, receptor dephosphorylation, and receptor recycling are required for resensitization, since inhibition of several of these steps suppresses resensitization of the NK1-R and the beta 2-AR.

Endocytosis and lysosomal degradation of PARs could contribute to desensitization by removing receptors from the cell surface and by degrading activated receptors in lysosomes. However, responses to PAR-2 desensitize when endocytosis is inhibited. This finding indicates that endocytosis is not the principal mechanism of attenuation (16).

Recovery of cellular responses to proteases requires the availability of intact receptors at the cell surface. Two general mechanisms account for this resensitization of PAR-1: synthesis of new receptors and mobilization of intracellular pools of receptors. The importance of these mechanisms depends on whether cells possess prominent intracellular stores of PAR-1. In megakaryoblastic HEL and CHRF-288 cells, where most receptors are rapidly internalized after thrombin cleavage, recovery is a slow process that depends on the synthesis of new receptors (22, 67). In contrast, endothelial cells and fibroblasts have large intracellular stores of PAR-1, and resensitization is rapid and associated with depletion of these pools and is initially independent of new receptor synthesis (60, 64). Similarly, resensitization of trypsin responses of transfected cells and enterocytes expressing PAR-2 at the cell surface and in prominent stores is minimally affected by cycloheximide but attenuated by disruption of the pools with brefeldin A, suggesting that resensitization requires mobilization of pools (17). Whether the mobilization of intracellular pools of PARs requires a specific signal or is a reflection of the tonic shuttling of receptors to and from the plasma membrane is unknown. Recovery of responses to multiple challenges with proteases is ultimately inhibited by cycloheximide, indicating that synthesis of new PARs becomes important once the pools are exhausted. In platelets, about two-thirds of PAR-1 is present at the plasma membrane, and the remainder is found in intracellular membranes (97). Interestingly, activation by ADP or thromboxane causes mobilization of these intracellular receptors to the plasma membrane, where they may be available for cleavage by thrombin. Because platelets lack the ability to synthesize new receptors and lack the intracellular pool found in endothelial cells, they are unable to repopulate the plasma membrane with new receptors after thrombin exposure and are thus unable to recover responsiveness to thrombin.

Endocytic and trafficking domains. Endocytosis and intracellular targeting of receptors require the interaction of specific receptor motifs with proteins that direct sorting (16). Endocytic motifs, which probably interact with proteins that direct clathrin-mediated endocytosis, have been identified in the intracellular COOH-tail of several single transmembrane domain receptors. These motifs are similar for different receptors and functionally interchangeable. Although endocytic domains have also been identified for several GPCRs, a common and interchangeable endocytic motif has not been identified, and it is unclear whether these motifs resemble those of the single transmembrane domain receptors.

The PARs may have several functionally important motifs that determine agonist-induced endocytosis and lysosomal targeting, constitutive agonist-independent shuttling between the plasma membrane and intracellular pools, and targeting of newly synthesized receptor to the intracellular stores. As yet, no specific, interchangeable motif has been identified, although some motifs have been characterized for PAR-1. PAR-1 activation is associated with receptor phosphorylation by GRKs and possibly PKC, which contribute to desensitization (72, 157, 158). Mutation of all the Ser and Thr residues to Ala in the COOH-tail abolishes agonist-induced phosphorylation and agonist-stimulated endocytosis of PAR-1, although tonic cycling of the receptor is unaffected (140). Therefore, receptor phosphorylation may be necessary for agonist-induced endocytosis but not tonic cycling of PAR-1. Truncation of PAR-1 at Tyr397 suppresses agonist-dependent and -independent internalization. A domain between K397SAMLQGSSNY407 that is critical for both forms of internalization was identified by analysis of a series of truncation mutants. Together, these results suggest that receptor phosphorylation is required for endocytosis. Phosphorylation of the beta 2-AR and muscarinic M2 receptor by GRK-2 is also important for agonist-induced endocytosis (47, 94, 130, 151). beta -Arrestins may interact with phosphorylated receptors and participate in endocytosis as well as uncoupling from G proteins. In support of this possibility, beta -arrestin binds to clathrin with high affinity, and beta 2-AR colocalizes with beta -arrestin and clathrin in the first-formed endosomes (46, 54). Thus beta -arrestins may serve as adaptor molecules that recruit cellular proteins that facilitate endocytosis of several GPCRs or that directly mediate endocytosis themselves.

Domains in the COOH-tail of PAR-1 also govern targeting to the Golgi apparatus in the absence of agonist. Although wild-type PAR-1 and PAR-1 with all COOH-terminal Ser and Thr residues mutated to Ala are expressed at the cell surface and in a large Golgi pool, the Tyr397 truncation mutant is only located at the cell surface (140).

The lysosomal targeting motifs of PAR-1 and PAR-2 have not been identified. Most of the information about lysosomal targeting derives from studies of lysosomal membrane proteins and growth factor receptors (16). Tyr-containing motifs in the COOH-tails are critical for lysosomal targeting of lysosome-associated membrane proteins I and II and lysosomal acid phosphatase. Lysosomal integral membrane protein II, which does not contain a Tyr residue in its COOH-tail, relies on a dileucine-related Leu-Ile motif for lysosomal targeting.

PAR Downregulation

Downregulation is characterized by a decrease in the total number of receptors in a cell and is caused by long-term exposure to agonists for hours or days; recovery from downregulation is similarly slow. From a physiological viewpoint, it is probably rare that a cell is continuously exposed to agonists, since efficient mechanisms exist to remove them from the extracellular fluid. However, downregulation may occur under pathological circumstances, when there is continuous production of an agonist, or during long-term administration of receptor agonists for therapeutic reasons. Although far less is known about receptor downregulation than desensitization, possible mechanisms include enhanced degradation and reduced synthesis of receptors.

The downregulation of PAR-1 has been examined in mesangial cells (166). Prolonged activation of PAR-1 results in reduced protein levels, whereas PAR-1 mRNA is unaffected, suggesting that increased degradation is the principal mechanism of receptor downregulation. Agonist-induced downregulation is not mediated by PKC because it is unaffected by PKC inhibitors. However, chronic activation of PKC with phorbol esters and of PKA with prostaglandin E1 diminishes PAR-1 protein and mRNA, suggesting that an inhibition of expression also contributes to downregulation. Together, these results indicate that there are distinct mechanisms for homologous and heterologous downregulation of PAR-1.

    CONCLUSIONS AND FUTURE PERSPECTIVES
Top
Abstract
Introduction
Conclusions
References

PARs comprise a new and growing family of GPCRs that mediate the biological actions of certain Ser proteases. Thus, in addition to their role as proteases that degrade extracellular proteins, Ser proteases function as signaling molecules that specifically regulate target cells by cleaving PARs. Thrombin is an established signaling protease that exerts many of its effects through PAR-1 and PAR-3. Trypsin may function as a signaling molecule in the lumen of the small intestine by cleaving and activating PAR-2 on enterocytes. However, the widespread distribution of PAR-2 suggests that different enzymes activate PAR-2 in other locations. To date the best candidate is mast cell tryptase, suggesting a role for PAR-2 in inflammation.

Comparisons between the mechanisms of signaling by GPCRs for proteases and neuropeptides reveal interesting similarities and distinct differences. PARs can be viewed as specialized peptide receptors, but ones in which the peptide ligand is physically part of the receptor molecule. Both types of receptors are activated by interaction of specific residues of the ligand with extracellular and transmembrane domains of the receptor. Proteases are important in initiating signaling of protease and neuropeptide receptors: cleavage activates PARs, and proteolysis is important for generation of biologically active peptides. Proteases are also important in terminating the signal: once cleaved, PARs can no longer be activated, and proteases degrade and inactivate neuropeptides. Receptors for proteases and neuropeptides couple to similar mechanisms of signal transduction, and signaling is efficiently attenuated by comparable processes of receptor desensitization. However, there are distinct differences in the intracellular trafficking of PARs and neuropeptide receptors: PARs are mostly degraded, but neuropeptide receptors recycle. Thus resensitization of responses to proteases requires mobilization of Golgi pools or synthesis of new receptor molecules.

There is much to be learned about PARs. New family members will probably emerge, as exemplified by the recent identification of PAR-3. The importance of PAR-1 is firmly established, but the role of PAR-2 and PAR-3 in physiology and pathophysiology is still emerging. Definitive studies of the importance of PAR-2 will depend on the availability of selective and potent antagonists and agonists and of animals in which the PAR-2 and PAR-3 genes are manipulated.

    ACKNOWLEDGEMENTS

We thank Dr. Katie DeFea for critically reading this manuscript.

    FOOTNOTES

This work was supported by National Institute of Diabetes and Digestive and Kidney Diseases Grants DK-43207 and DK-39957.

O. Déry and C. U. Covera contributed equally to this work.

Address for reprint requests: N. Bunnett, Box 0660, Univ. of California, San Francisco, 521 Parnassus Ave., San Francisco, CA 94143-0660.

    REFERENCES
Top
Abstract
Introduction
Conclusions
References

1.   Al-Ani, B., M. Saifeddine, and M. D. Hollenberg. 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, 1995[Medline].

2.   Albrightson, C. R., B. Zabko-Potapovich, G. Dytko, W. M. Bryan, K. Hoyle, M. L. Moore, and J. M. Stadel. Analogues of the thrombin receptor tethered-ligand enhance mesangial cell proliferation. Cell. Signal. 6: 743-750, 1994[Medline].

3.   Antonaccio, M. J., D. Normandin, R. Serafino, and S. Moreland. Effects of thrombin and thrombin receptor activating peptides on rat aortic vascular smooth muscle. J. Pharmacol. Exp. Ther. 266: 125-132, 1993[Abstract].

4.   Aragay, A. M., L. R. Collins, G. R. Post, A. J. Watson, J. R. Feramisco, J. H. Brown, and M. I. Simon. G12 requirement for thrombin-stimulated gene expression and DNA synthesis in 1321N1 astrocytoma cells. J. Biol. Chem. 270: 20073-20077, 1995[Abstract/Free Full Text].

5.   Arena, C. S., S. M. Quirk, Y. Q. Zhang, and K. P. Henrikson. Rat uterine stromal cells: thrombin receptor and growth stimulation by thrombin. Endocrinology 137: 3744-3749, 1996[Abstract].

6.   Austyn, J., and K. Wood. Principles of Cellular and Molecular Immunology. Oxford, UK: Oxford Univ. Press, 1993, p. 499-580.

7.   Baffy, G., L. Yang, S. Raj, D. R. Manning, and J. R. Williamson. G protein coupling to the thrombin receptor in Chinese hamster lung fibroblasts. J. Biol. Chem. 269: 8483-8487, 1994[Abstract/Free Full Text].

8.   Bar-Shavit, R., M. Benezra, V. Sabbah, W. Bode, and I. Vlodavsky. Thrombin as a multifunctional protein: induction of cell adhesion and proliferation. Am. J. Respir. Cell Mol. Biol. 6: 123-130, 1992[Medline].

9.   Beecher, K. L., T. T. Andersen, J. W. Fenton II, and B. W. Festoff. Thrombin receptor peptides induce shape change in neonatal murine astrocytes in culture. J. Neurosci. Res. 37: 108-115, 1994[Medline].

10.   Belham, C. M., R. J. Tate, P. H. Scott, A. D. Pemberton, H. R. Miller, R. M. Wadsworth, G. W. Gould, and R. Plevin. Trypsin stimulates proteinase-activated receptor-2-dependent and -independent activation of mitogen-activated protein kinases. Biochem. J. 320: 939-946, 1996[Medline].

11.   Benka, M. L., M. Lee, G.-R. Wang, S. Buckman, A. Burlacu, L. Cole, A. DePina, P. Dias, A. Granger, B. Grant, The thrombin receptor in human platelets is coupled to a GTP binding protein of the Galpha q family. FEBS Lett. 363: 49-52, 1995[Medline].

12.   Bernatowicz, M. S., C. E. Klimas, K. S. Hartl, M. Peluso, N. J. Allegretto, and S. M. Seiler. Development of potent thrombin receptor antagonist peptides. J. Med. Chem. 39: 4879-4887, 1996[Medline].

13.   Blackhart, B. D., K. Emilsson, D. Nguyen, W. Teng, A. J. Martelli, S. Nystedt, J. Sundelin, and R. M. Scarborough. Ligand cross-reactivity within the protease-activated receptor family. J. Biol. Chem. 271: 16466-16471, 1996[Abstract/Free Full Text].

14.   Blenis, J. Signal transduction via the MAP kinases: proceed at your own RSK. Proc. Natl. Acad. Sci. USA 90: 5889-5892, 1993[Abstract].

16.   Böhm, S., E. F. Grady, and N. W. Bunnett. Mechanisms attenuating signaling by G-protein coupled receptors. Biochem. J. 322: 1-18, 1997[Medline].

17.   Böhm, S. K., L. M. Khitin, E. F. Grady, G. Aponte, D. G. Payan, and N. W. Bunnett. Mechanisms of desensitization and resensitization of proteinase-activated receptor-2. J. Biol. Chem. 271: 22003-22016, 1996[Abstract/Free Full Text].

18.   Böhm, S. K., W. Kong, D. Brömme, S. P. Smeekens, D. C. Anderson, A. Connolly, M. Kahn, N. A. Nelken, S. R. Coughlin, D. G. Payan, and N. W. Bunnett. Molecular cloning, expression and potential functions of the human proteinase-activated receptor-2. Biochem. J. 314: 1009-1016, 1996[Medline].

19.   Bouton, M. C., M. Jandrot-Perrus, S. Moog, J. P. Cazenave, M. C. Guillin, and F. Lanza. Thrombin interaction with a recombinant N-terminal extracellular domain of the thrombin receptor in an acellular system. Biochem. J. 305: 635-641, 1995[Medline].

20.   Boyd, N. D., R. Kage, J. J. Dumas, J. E. Krause, and S. E. Leeman. The peptide binding site of the substance P (NK-1) receptor localized by a photoreactive analogue of substance P: presence of a disulfide bond. Proc. Natl. Acad. Sci. USA 93: 433-437, 1996[Abstract/Free Full Text].

21.   Brass, L. F., and M. Molino. Protease-activated G protein-coupled receptors on human platelets and endothelial cells. Thromb. Haemost. 78: 234-241, 1997[Medline].

22.   Brass, L. F., S. Pizarro, M. Ahuja, E. Belmonte, N. Blanchard, J. M. Stadel, and J. A. Hoxie. Changes in the structure and function of the human thrombin receptor during receptor activation, internalization, and recycling. J. Biol. Chem. 269: 2943-2952, 1994[Abstract/Free Full Text].

23.   Brown, J. K., C. A. Jones, C. L. Tyler, S. J. Ruoss, T. Hartmann, and G. H. Caughey. Tryptase-induced mitogenesis in airway smooth muscle cells. Potency, mechanisms, and interactions with other mast cell mediators. Chest 107: 95S-96S, 1995[Medline].

24.   Brown, J. K., C. L. Tyler, C. A. Jones, S. J. Ruoss, T. Hartmann, and G. H. Caughey. Tryptase, the dominant secretory granular protein in human mast cells, is a potent mitogen for cultured dog tracheal smooth muscle cells. J. Clin. Invest. 88: 493-499, 1991[Medline].

25.   Cairns, J. A., and A. F. Walls. Mast cell tryptase is a mitogen for epithelial cells. Stimulation of IL-8 production and intracellular adhesion molecule-1 expression. J. Immunol. 156: 275-283, 1995[Abstract].

26.   Cavanaugh, K. P., D. Gurwitz, D. D. Cunningham, and R. A. Bradshaw. Reciprocal modulation of astrocyte stellation by thrombin and protease nexin-1. J. Neurochem. 54: 1735-1743, 1990[Medline].

27.   Chen, H. C., and J. L. Guan. Association of focal adhesion kinase with its potential substrate phosphatidylinositol 3-kinase. Proc. Natl. Acad. Sci. USA 91: 10148-10152, 1994[Abstract/Free Full Text].

28.   Chen, J., M. Ishii, L. Wang, K. Ishii, and S. R. Coughlin. Thrombin receptor activation. Confirmation of the intramolecular tethered liganding hypothesis and discovery of an alternative intermolecular liganding mode. J. Biol. Chem. 269: 16041-16045, 1994[Abstract/Free Full Text].

29.   Chen, X., K. Earley, W. Luo, S. H. Lin, and W. P. Schilling. Functional expression of a human thrombin receptor in Sf9 insect cells: evidence for an active tethered ligand. Biochem. J. 314: 603-611, 1996[Medline].

30.   Chen, Y., D. Grall, A. E. Salcini, P. G. Pelicci, J. Pouyssegur, and E. Van Obberghen-Schilling. Shc adaptor proteins are key transducers of mitogenic signaling mediated by the G protein-coupled thrombin receptor. EMBO J. 15: 1037-1044, 1996[Abstract].

31.   Chen, Y. H., J. Pouyssegur, S. A. Courtneidge, and E. Van Obberghen-Schilling. Activation of Src family kinase activity by the G protein-coupled thrombin receptor in growth-responsive fibroblasts. J. Biol. Chem. 269: 27372-27377, 1994[Abstract/Free Full Text].

32.   Cirino, G., C. Cicala, M. R. Bucci, L. Sorrentino, J. M. Maraganore, and S. R. Stone. Thrombin functions as an inflammatory mediator through activation of its receptor. J. Exp. Med. 183: 821-827, 1996[Abstract].

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

34.   Connolly, A. J., H. Ishihara, M. L. Kahn, R. V. Farese, Jr., and S. R. Coughlin. Role of the thrombin receptor in development and evidence for a second receptor. Nature 381: 516-519, 1996[Medline].

35.   Corvera, C. U., O. Déry, K. McConalogue, S. K. Böhm, L. M. Khitin, G. H. Caughey, D. G. Payan, and N. W. Bunnett. Mast cell tryptase regulates colonic myocytes through proteinase-activated receptor-2. J. Clin. Invest. 100: 1383-1393, 1997[Abstract/Free Full Text].

36.   Coughlin, S. R. Protease-activated receptors start a family. Proc. Natl. Acad. Sci. USA 91: 9200-9202, 1994[Free Full Text].

37.   Damiano, B. P., W. M. Cheung, J. A. Mitchell, and R. Falotico. Cardiovascular actions of thrombin receptor activation in vivo. J. Pharmacol. Exp. Ther. 279: 1365-1378, 1996[Abstract].

38.   Darrow, A. L., W. P. Fung-Leung, R. D. Ye, R. J. Santulli, W. M. Cheung, C. K. Derian, C. L. Burns, B. P. Damiano, L. Zhou, C. M. Keenan, P. A. Peterson, and P. Andrade-Gordon. Biological consequences of thrombin receptor deficiency in mice. Thromb. Haemost. 76: 860-866, 1996[Medline].

39.   DeBlois, D., G. Drapeau, E. Petitclerc, and F. Marceau. Synergism between the contractile effect of epidermal growth factor and that of des-Arg9-bradykinin or of alpha -thrombin in rabbit aortic rings. Br. J. Pharmacol. 105: 959-967, 1992[Abstract].

40.   Derian, C. K., A. J. Eckardt, and P. Andrade-Gordon. Differential regulation of human keratinocyte growth and differentiation by a novel family of protease-activated receptors. Cell Growth Differ. 8: 743-749, 1997[Abstract].

41.   Dihanich, M., M. Kaser, E. Reinhard, D. Cunningham, and D. Monard. Prothrombin mRNA is expressed by cells of the nervous system. Neuron 6: 575-581, 1991[Medline].

42.   Ehrenreich, H., T. Costa, K. A. Clouse, R. M. Pluta, Y. Ogino, J. E. Coligan, and P. R. Burd. Thrombin is a regulator of astrocytic endothelin-1. Brain Res. 600: 201-207, 1993[Medline].

43.   Elling, C. E., S. M. Nielsen, and T. W. Schwartz. Conversion of antagonist-binding site to metal-ion site in the tachykinin NK-1 receptor. Nature 374: 74-77, 1995[Medline].

44.   Exton, J. H. Phosphoinositide phospholipases and G proteins in hormone action. Annu. Rev. Physiol. 56: 349-369, 1994[Medline].

45.   Fee, J. A., J. D. Monsey, R. J. Handler, M. A. Leonis, S. R. Mullaney, H. M. Hope, and D. F. Silbert. A Chinese hamster fibroblast mutant defective in thrombin-induced signaling has a low level of phospholipase C-beta 1. J. Biol. Chem. 269: 21699-21708, 1994[Abstract/Free Full Text].

46.   Ferguson, S. S., W. E. Downey, A. M. Colapietro, L. S. Barak, L. Menard, and M. G. Caron. Role of beta -arrestin in mediating agonist-promoted G protein-coupled receptor internalization. Science 271: 363-366, 1996[Abstract].

47.   Ferguson, S. S., L. Menard, L. S. Barak, W. J. Koch, A. M. Colapietro, and M. G. Caron. Role of phosphorylation in agonist-promoted beta 2-adrenergic receptor sequestration. Rescue of a sequestration-defective mutant receptor by beta  ARK1. J. Biol. Chem. 270: 24782-24789, 1995[Abstract/Free Full Text].

48.   Festoff, B. W., I. V. Smirnova, J. Ma, and B. A. Citron. Thrombin, its receptor and protease nexin I, its potent serpin, in the nervous system. Semin. Thromb. Hemost. 22: 267-271, 1996[Medline].

49.   Fischer, T. H., M. N. Gatling, J. C. Lacal, and G. C. White II. rap1B, a cAMP-dependent protein kinase substrate, associates with the platelet cytoskeleton. J. Biol. Chem. 265: 19405-19408, 1990[Abstract/Free Full Text].

50.   Gerszten, R. E., J. Chen, M. Ishii, K. Ishii, L. Wang, T. Nanevicz, C. W. Turck, T. K. Vu, and S. R. Coughlin. Specificity of the thrombin receptor for agonist peptide is defined by its extracellular surface. Nature 368: 648-651, 1994[Medline].

51.   Gether, U., T. E. Johansen, and T. W. Schwartz. Chimeric NK1 (substance P)/NK3 (neurokinin B) receptors. Identification of domains determining the binding specificity of tachykinin agonists. J. Biol. Chem. 268: 7893-7898, 1993[Abstract/Free Full Text].

52.   Gether, U., L. Nilsson, J. A. Lowe III, and T. W. Schwartz. Specific residues at the top of transmembrane segment V and VI of the neurokinin-1 receptor involved in binding of the nonpeptide antagonist CP 96,345. J. Biol. Chem. 269: 23959-23964, 1994[Abstract/Free Full Text] 269: December 1994, p. 32708.]

53.   Godin, D., F. Marceau, C. Beaule, F. Rioux, and G. Drapeau. Aminopeptidase modulation of the pharmacological responses to synthetic thrombin receptor agonists. Eur. J. Pharmacol. 253: 225-230, 1994[Medline].

54.   Goodman, O. B., Jr., J. G. Krupnick, F. Santini, V. V. Gurevich, R. B. Penn, A. W. Gagnon, J. H. Keen, and J. L. Benovic. beta -Arrestin acts as a clathrin adaptor in endocytosis of the beta 2-adrenergic receptor. Nature 383: 447-450, 1996[Medline].

55.   Grand, R. J., A. S. Turnell, and P. W. Grabham. Cellular consequences of thrombin-receptor activation. Biochem. J. 313: 353-368, 1996[Medline].

56.   Grandaliano, G., A. J. Valente, and H. E. Abboud. A novel biologic activity of thrombin: stimulation of monocyte chemotactic protein production. J. Exp. Med. 179: 1737-1741, 1994[Abstract].

57.   Guinebault, C., B. Payrastre, C. Sultan, G. Mauco, M. Breton, S. Levy-Toledano, M. Plantavid, and H. Chap. Tyrosine kinases and phosphoinositide metabolism in thrombin-stimulated human platelets. Biochem. J. 292: 851-856, 1993[Medline].

58.   Gurwitz, D., and D. D. Cunningham. Thrombin modulates and reverses neuroblastoma neurite outgrowth. Proc. Natl. Acad. Sci. USA 85: 3440-3444, 1988[Abstract].

59.   Hartmann, T., S. J. Ruoss, W. W. Raymond, K. Seuwen, and G. H. Caughey. Human tryptase as a potent, cell-specific mitogen: role of signaling pathways in synergistic responses. Am. J. Physiol. 262 (Lung Cell. Mol. Physiol. 6): L528-L534, 1992[Abstract/Free Full Text].

60.   Hein, L., K. Ishii, S. R. Coughlin, and B. K. Kobilka. Intracellular targeting and trafficking of thrombin receptors. A novel mechanism for resensitization of a G protein-coupled receptor. J. Biol. Chem. 269: 27719-27726, 1994[Abstract/Free Full Text].

61.   Hollenberg, M. D. Protease-mediated signalling: new paradigms for cell regulation and drug development. Trends Pharmacol. Sci. 17: 3-6, 1996[Medline].

62.   Hollenberg, M. D., A. A. Laniyonu, M. Saifeddine, and G. J. Moore. Role of the amino- and carboxyl-terminal domains of thrombin receptor-derived polypeptides in biological activity in vascular endothelium and gastric smooth muscle: evidence for receptor subtypes. Mol. Pharmacol. 43: 921-930, 1993[Abstract].

63.   Hollenberg, M. D., M. Saifeddine, and B. Al-Ani. Proteinase-activated receptor-2 in rat aorta: structural requirements for agonist activity of receptor-activating peptides. Mol. Pharmacol. 49: 229-233, 1996[Abstract].

64.   Horvat, R., and G. E. Palade. The functional thrombin receptor is associated with the plasmalemma and a large endosomal network in cultured human umbilical vein endothelial cells. J. Cell Sci. 108: 1155-1164, 1995[Abstract/Free Full Text].

65.   Howells, G. L., M. Macey, M. A. Curtis, and S. R. Stone. Peripheral blood lymphocytes express the platelet-type thrombin receptor. Br. J. Haematol. 84: 156-160, 1993[Medline].

66.   Howells, G. L., M. G. Macey, C. Chinni, L. Hou, M. T. Fox, P. Harriott, and S. R. Stone. Proteinase-activated receptor-2: expression by human neutrophils. J. Cell Sci. 110: 881-887, 1997[Abstract/Free Full Text].

67.   Hoxie, J. A., M. Ahuja, E. Belmonte, S. Pizarro, R. Parton, and L. F. Brass. Internalization and recycling of activated thrombin receptors. J. Biol. Chem. 268: 13756-13763, 1993[Abstract/Free Full Text].

68.   Hung, D. T., T. K. Vu, V. I. Wheaton, K. Ishii, and S. R. Coughlin. Cloned platelet thrombin receptor is necessary for thrombin-induced platelet activation. J. Clin. Invest. 89: 1350-1353, 1992[Medline].

69.   Hung, D. T., Y. H. Wong, T. K. Vu, and S. R. Coughlin. 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, 1992[Abstract/Free Full Text].

70.   Hwa, J. J., L. Ghibaudi, P. Williams, M. Chintala, R. Zhang, M. Chatterjee, and E. Sybertz. Evidence for the presence of a proteinase-activated receptor distinct from the thrombin receptor in vascular endothelial cells. Circ. Res. 78: 581-588, 1996[Abstract/Free Full Text].

71.   Ishihara, H., A. J. Connolly, D. Zeng, M. L. Kahn, Y. W. Zheng, C. Timmons, T. Tram, and S. R. Coughlin. Protease-activated receptor 3 is a second thrombin receptor in humans. Nature 386: 502-506, 1997[Medline].

72.   Ishii, K., J. Chen, M. Ishii, W. J. Koch, N. J. Freedman, R. J. Lefkowitz, and S. R. Coughlin. Inhibition of thrombin receptor signaling by a G-protein coupled receptor kinase. Functional specificity among G-protein coupled receptor kinases. J. Biol. Chem. 269: 1125-1130, 1994[Abstract/Free Full Text].

73.   Ishii, K., R. Gerszten, Y. W. Zheng, J. B. Welsh, C. W. Turck, and S. R. Coughlin. Determinants of thrombin receptor cleavage. Receptor domains involved, specificity, and role of the P3 aspartate. J. Biol. Chem. 270: 16435-16440, 1995[Abstract/Free Full Text].

74.   Ishii, K., L. Hein, B. Kobilka, and S. R. Coughlin. Kinetics of thrombin receptor cleavage on intact cells. Relation to signaling. J. Biol. Chem. 268: 9780-9786, 1993[Abstract/Free Full Text].

75.   Jalink, K., E. J. van Corven, T. Hengeveld, N. Morii, S. Narumiya, and W. H. Moolenaar. Inhibition of lysophosphatidate- and thrombin-induced neurite retraction and neuronal cell rounding by ADP ribosylation of the small GTP-binding protein Rho. J. Cell Biol. 126: 801-810, 1994[Abstract].

76.   Jenkins, A. L., M. D. Bootman, C. W. Taylor, E. J. Mackie, and S. R. Stone. Characterization of the receptor responsible for thrombin-induced intracellular calcium responses in osteoblast-like cells. J. Biol. Chem. 268: 21432-21437, 1993[Abstract/Free Full Text].

77.   Jones, A., and C. L. Geczy. Thrombin and factor Xa enhance the production of interleukin-1. Immunology 71: 236-241, 1990[Medline].

78.   Kanthou, C., O. Benzakour, G. Patel, J. Deadman, V. V. Kakkar, and F. Lupu. Thrombin receptor activating peptide (TRAP) stimulates mitogenesis, c-fos and PDGF-A gene expression in human vascular smooth muscle cells. Thromb. Haemost. 74: 1340-1347, 1995[Medline].

79.   King, W. G., G. L. Kucera, A. Sorisky, J. Zhang, and S. E. Rittenhouse. Protein kinase C regulates the stimulated accumulation of 3-phosphorylated phosphoinositides in platelets. Biochem. J. 278: 475-480, 1991[Medline].

80.   Kinlough-Rathbone, R. L., M. L. Rand, and M. A. Packham. Rabbit and rat platelets do not respond to thrombin receptor peptides that activate human platelets. Blood 82: 103-106, 1993[Abstract].

81.   Klepfish, A., M. A. Greco, and S. Karpatkin. Thrombin stimulates melanoma tumor-cell binding to endothelial cells and subendothelial matrix. Int. J. Cancer 53: 978-982, 1993[Medline].

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

83.   Kong, W., K. McConalogue, L. M. Khitin, M. D. Hollenberg, D. G. Payan, S. K. Bohm, and N. W. Bunnett. Luminal trypsin may regulate enterocytes through proteinase-activated receptor 2. Proc. Natl. Acad. Sci. USA 94: 8884-8889, 1997[Abstract/Free Full Text].

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

85.   Ku, D. D. Mechanism of thrombin-induced endothelium-dependent coronary vasodilation in dogs: role of its proteolytic enzymatic activity. J. Cardiovasc. Pharmacol. 8: 29-36, 1986[Medline].

86.   L'Allemain, G., J. Pouyssegur, and M. J. Weber. p42/Mitogen-activated protein kinase as a converging target for different growth factor signaling pathways: use of pertussis toxin as a discrimination factor. Cell Regul. 2: 675-684, 1991[Medline].

87.   Laposata, M., D. K. Dovnarsky, and H. S. Shin. Thrombin-induced gap formation in confluent endothelial cell monolayers in vitro. Blood 62: 549-556, 1983[Abstract].

88.   Lau, L. F., K. Pumiglia, Y. P. Cote, and M. B. Feinstein. Thrombin-receptor agonist peptides, in contrast to thrombin itself, are not full agonists for activation and signal transduction in human platelets in the absence of platelet-derived secondary mediators. Biochem. J. 303: 391-400, 1994[Medline].

89.   Lerner, D. J., M. Chen, T. Tram, and S. R. Coughlin. Agonist recognition by proteinase-activated receptor 2 and thrombin receptor. Importance of extracellular loop interactions for receptor function. J. Biol. Chem. 271: 13943-13947, 1996[Abstract/Free Full Text].

90.   Lum, H., T. T. Andersen, A. Siflinger-Birnboim, C. Tiruppathi, M. S. Goligorsky, J. W. Fenton II, and A. B. Malik. Thrombin receptor peptide inhibits thrombin-induced increase in endothelial permeability by receptor desensitization. J. Cell Biol. 120: 1491-1499, 1993[Abstract].

91.   Lum, H., J. L. Aschner, P. G. Phillips, P. W. Fletcher, and A. B. Malik. Time course of thrombin-induced increase in endothelial permeability: relationship to Ca2+i and inositol polyphosphates. Am. J. Physiol. 263 (Lung Cell. Mol. Physiol. 7): L219-L225, 1992[Abstract/Free Full Text]. 7): October 1992, following table of contents.]

92.   Mari, B., S. Guerin, D. F. Far, J. P. Breitmayer, N. Belhacene, J. F. Peyron, B. Rossi, and P. Auberger. Thrombin and trypsin-induced Ca(2+) mobilization in human T cell lines through interaction with different protease-activated receptors. FASEB J. 10: 309-316, 1996[Abstract/Free Full Text].

93.   McNamara, C. A., I. J. Sarembock, L. W. Gimple, J. W. Fenton II, S. R. Coughlin, and G. K. Owens. Thrombin stimulates proliferation of cultured rat aortic smooth muscle cells by a proteolytically activated receptor. J. Clin. Invest. 91: 94-98, 1993[Medline].

94.   Menard, L., S. S. Ferguson, L. S. Barak, L. Bertrand, R. T. Premont, A. M. Colapietro, R. J. Lefkowitz, and M. G. Caron. Members of the G protein-coupled receptor kinase family that phosphorylate the beta 2-adrenergic receptor facilitate sequestration. Biochemistry 35: 4155-4160, 1996[Medline].

95.   Miller, S., N. Sehati, C. Romano, and C. W. Cotman. Exposure of astrocytes to thrombin reduces levels of the metabotropic glutamate receptor mGluR5. J. Neurochem. 67: 1435-1447, 1996[Medline].

96.   Mirza, H., V. Yatsula, and W. F. Bahou. The proteinase activated receptor-2 (PAR-2) mediates mitogenic responses in human vascular endothelial cells. J. Clin. Invest. 97: 1705-1714, 1996[Abstract/Free Full Text].

97.   Molino, M., D. F. Bainton, J. A. Hoxie, S. R. Coughlin, and L. F. Brass. Thrombin receptors on human platelets. Initial localization and subsequent redistribution during platelet activation. J. Biol. Chem. 272: 6011-6017, 1997[Abstract/Free Full Text].

98.   Molino, M., E. S. Barnathan, R. Numerof, J. Clark, M. Dreyer, A. Cumashi, J. Hoxie, N. Schechter, M. Woolkalis, and L. F. Brass. Interactions of mast cell tryptase with thrombin receptors and PAR-2. J. Biol. Chem. 272: 4043-4049, 1997[Abstract/Free Full Text].

99.   Molino, M., N. Blanchard, E. Belmonte, A. P. Tarver, C. Abrams, J. A. Hoxie, C. Cerletti, and L. F. Brass. Proteolysis of the human platelet and endothelial cell thrombin receptor by neutrophil-derived cathepsin G. J. Biol. Chem. 270: 11168-11175, 1995[Abstract/Free Full Text].

100.   Molloy, C. J., J. E. Pawlowski, D. S. Taylor, C. E. Turner, H. Weber, and M. Peluso. Thrombin receptor activation elicits rapid protein tyrosine phosphorylation and stimulation of the raf-1/MAP kinase pathway preceding delayed mitogenesis in cultured rat aortic smooth muscle cells: evidence for an obligate autocrine mechanism promoting cell proliferation induced by G-protein-coupled receptor agonist. J. Clin. Invest. 97: 1173-1183, 1996[Abstract/Free Full Text].

101.   Morris, R., P. G. Winyard, D. R. Blake, and C. J. Morris. Thrombin in inflammation and healing: relevance to rheumatoid arthritis. Ann. Rheum. Dis. 53: 72-79, 1994[Medline].

102.   Morris, R., P. G. Winyard, L. F. Brass, D. R. Blake, and C. J. Morris. Thrombin receptor expression in rheumatoid and osteoarthritic synovial tissue. Ann. Rheum. Dis. 55: 841-843, 1996[Abstract].

103.   Moser, R., P. Groscurth, and J. Fehr. Promotion of transendothelial neutrophil passage by human thrombin. J. Cell Sci. 96: 737-744, 1990[Abstract].

104.   Muramatsu, I., A. Laniyonu, G. J. Moore, and M. D. Hollenberg. Vascular actions of thrombin receptor peptide. Can. J. Physiol. Pharmacol. 70: 996-1003, 1992[Medline].

105.   Nanevicz, T., M. Ishii, L. Wang, M. Chen, J. Chen, C. W. Turck, F. E. Cohen, and S. R. Coughlin. Mechanisms of thrombin receptor agonist specificity. Chimeric receptors and complementary mutations identify an agonist recognition site. J. Biol. Chem. 270: 21619-21625, 1995[Abstract/Free Full Text].

106.   Nanevicz, T., L. Wang, M. Chen, M. Ishii, and S. R. Coughlin. Thrombin receptor activating mutations. Alteration of an extracellular agonist recognition domain causes constitutive signaling. J. Biol. Chem. 271: 702-706, 1996[Abstract/Free Full Text].

107.   Nawa, H., H. Kotani, and S. Nakanishi. Tissue-specific generation of two preprotachykinin mRNAs from one gene by alternative RNA splicing. Nature 312: 729-734, 1984[Medline].

108.   Nelken, N. A., S. J. Soifer, J. O'Keefe, T. K. Vu, I. F. Charo, and S. R. Coughlin. Thrombin receptor expression in normal and atherosclerotic human arteries. J. Clin. Invest. 90: 1614-1621, 1992[Medline].

109.   Neveu, I., F. Jehan, M. Jandrot-Perrus, D. Wion, and P. Brachet. Enhancement of the synthesis and secretion of nerve growth factor in primary cultures of glial cells by proteases: a possible involvement of thrombin. J. Neurochem. 60: 858-867, 1993[Medline].

111.   Niclou, S., H. S. Suidan, M. Brown-Luedi, and D. Monard. Expression of the thrombin receptor mRNA in rat brain. Cell. Mol. Biol. (Noisy-le-grand) 40: 421-428, 1994.

112.   Nierodzik, M. L., R. M. Bain, L. X. Liu, M. Shivji, K. Takeshita, and S. Karpatkin. Presence of the seven transmembrane thrombin receptor on human tumour cells: effect of activation on tumour adhesion to platelets and tumor tyrosine phosphorylation. Br. J. Haematol. 92: 452-457, 1996[Medline].

113.   Nierodzik, M. L., F. Kajumo, and S. Karpatkin. Effect of thrombin treatment of tumor cells on adhesion of tumor cells to platelets in vitro and tumor metastasis in vivo. Cancer Res. 52: 3267-3272, 1992[Abstract].

114.   Norton, K. J., R. M. Scarborough, J. L. Kutok, M. A. Escobedo, L. Nannizzi, and B. S. Coller. Immunologic analysis of the cloned platelet thrombin receptor activation mechanism: evidence supporting receptor cleavage, release of the N-terminal peptide, and insertion of the tethered ligand into a protected environment. Blood 82: 2125-2136, 1993[Abstract].

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

116.   Nystedt, S., K. Emilsson, C. Wahlestedt, and J. Sundelin. Molecular cloning of a potential proteinase activated receptor. Proc. Natl. Acad. Sci. USA 91: 9208-9212, 1994[Abstract/Free Full Text].

117.   Nystedt, S., A. K. Larsson, H. Aberg, and J. Sundelin. The mouse proteinase-activated receptor-2 cDNA and gene. Molecular cloning and functional expression. J. Biol. Chem. 270: 5950-5955, 1995[Abstract/Free Full Text].

118.   Nystedt, S., V. Ramakrishnan, and J. Sundelin. 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, 1996[Abstract/Free Full Text].

119.   Offermanns, S., K. L. Laugwitz, K. Spicher, and G. Schultz. 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, 1994[Abstract].

120.   Otsuka, M., and K. Yoshioka. Neurotransmitter functions of mammalian tachykinins. Physiol. Rev. 73: 229-308, 1993[Free Full Text].

121.   Papkoff, J., R. H. Chen, J. Blenis, and J. Forsman. p42 Mitogen-activated protein kinase and p90 ribosomal S6 kinase are selectively phosphorylated and activated during thrombin-induced platelet activation and aggregation. Mol. Cell. Biol. 14: 463-472, 1994[Abstract].

122.   Parma, J., L. Duprez, J. V. Sande, P. Cochaux, C. Gervy, J. Mockel, J. Dumont, and G. Vassarg. Somatic mutations in the thyrotropin receptor gene cause hyperfunctioning thyroid adenomas. Nature 365: 649-651, 1993[Medline].

123.   Perraud, F., F. Besnard, G. Labourdette, and M. Sensenbrenner. Proliferation of rat astrocytes, but not of oligodendrocytes, is stimulated in vitro by protease inhibitors. Int. J. Dev. Neurosci. 6: 261-266, 1988[Medline].

124.   Perraud, F., F. Besnard, M. Sensenbrenner, and G. Labourdette. Thrombin is a potent mitogen for rat astroblasts but not for oligodendroblasts and neuroblasts in primary culture. Int. J. Dev. Neurosci. 5: 181-188, 1987[Medline].

125.   Pike, C. J., P. J. Vaughan, D. D. Cunningham, and C. W. Cotman. Thrombin attenuates neuronal cell death and modulates astrocyte reactivity induced by beta -amyloid in vitro. J. Neurochem. 66: 1374-1382, 1996[Medline].

126.   Post, G. R., L. R. Collins, E. D. Kennedy, S. A. Moskowitz, A. M. Aragay, D. Goldstein, and J. H. Brown. 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, 1996[Abstract].

127.   Razin, E., and G. Marx. Thrombin-induced degranulation of cultured bone marrow-derived mast cells. J. Immunol. 133: 3282-3285, 1984[Abstract/Free Full Text].

128.   Renesto, P., M. Si-Tahar, M. Moniatte, V. Balloy, A. Van Dorsselaer, D. Pidard, and M. Chignard. Specific inhibition of thrombin-induced cell activation by the neutrophil proteinases elastase, cathepsin G, and proteinase 3: evidence for distinct cleavage sites within the aminoterminal domain of the thrombin receptor. Blood 89: 1944-1953, 1997[Abstract/Free Full Text].

129.   Roques, B. P., F. Noble, V. Dauge, M. C. Fournie-Zaluski, and A. Beaumont. Neutral endopeptidase 24.11: structure, inhibition, and experimental and clinical pharmacology. Pharmacol. Rev. 45: 87-146, 1993[Medline].

130.   Ruiz-Gomez, A., and F. Mayor. beta -Adrenergic receptor kinase (GRK2) colocalizes with beta -adrenergic receptors during agonist-induced receptor internalization. J. Biol. Chem. 272: 9601-9604, 1997[Abstract/Free Full Text].

131.   Ruoss, S., T. Hartmann, and G. Caughey. Mast cell tryptase is a mitogen for cultured fibroblasts. J. Clin. Invest. 88: 493-499, 1991[Medline].

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

133.   Santulli, R. J., C. K. Derian, A. L. Darrow, K. A. Tomko, A. J. Eckardt, M. Seiberg, R. M. Scarborough, and P. Andrade-Gordon. 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, 1995[Abstract].

134.   Satoh, K., Y. Ozaki, N. Asazuma, Y. Yatomi, Q. Ruomei, K. Kuroda, L. Yang, and S. Kume. Differential mobilization of tyrosine kinases in human platelets stimulated with thrombin or thrombin receptor agonist peptide. Biochem. Biophys. Res. Commun. 225: 1084-1089, 1996[Medline].

135.   Scarborough, R. M., M. A. Naughton, W. Teng, D. T. Hung, J. Rose, T. K. Vu, V. I. Wheaton, C. W. Turck, and S. R. Coughlin. Tethered ligand agonist peptides. Structural requirements for thrombin receptor activation reveal mechanism of proteolytic unmasking of agonist function. J. Biol. Chem. 267: 13146-13149, 1992[Abstract/Free Full Text].

136.   Schaeffer, P., E. Riera, E. Dupuy, and J. M. Herbert. 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, 1997[Medline].

137.   Schmidt, V. A., W. C. Nierman, T. V. Feldblyum, D. R. Maglott, and W. F. Bahou. The human thrombin receptor and proteinase activated receptor-2 genes are tightly linked on chromosome 5q13. Br. J. Haematol. 97: 523-529, 1997[Medline].

138.   Schmidt, V. A., E. Vitale, and W. F. Bahou. 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, 1996[Abstract/Free Full Text].

139.   Seiler, S. M., M. Peluso, I. M. Michel, H. Goldenberg, J. W. Fenton II, D. Riexinger, and S. Natarajan. Inhibition of thrombin and SFLLR-peptide stimulation of platelet aggregation, phospholipase A2 and Na+/H+ exchange by a thrombin receptor antagonist. Biochem. Pharmacol. 49: 519-528, 1995[Medline].

140.   Shapiro, M. J., J. Trejo, D. Zeng, and S. R. Coughlin. 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, 1996[Abstract/Free Full Text].

141.   Shenker, S., L. Laue, S. Kosugi, J. J. Merendino, T. Minegishi, and G. B. Cutler. A constitutively activating mutation of the luteinizing hormone receptor in familial male precocious puberty. Nature 365: 652-654, 1993[Medline].

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

143.   Shock, D. D., K. He, J. D. Wencel-Drake, and L. V. Parise. Ras activation in platelets after stimulation of the thrombin receptor, thromboxane A2 receptor or protein kinase C. Biochem. J. 321: 525-530, 1997[Medline].

144.   Simon, M., M. Strathmann, and N. Gautam. Diversity of G proteins in signal transduction. Science 252: 802-808, 1991[Medline].

145.   Smith-Swintosky, V. L., S. Zimmer, J. W. Fenton II, and M. P. Mattson. Opposing actions of thrombin and protease nexin-1 on amyloid beta -peptide toxicity and on accumulation of peroxides and calcium in hippocampal neurons. J. Neurochem. 65: 1415-1418, 1995[Medline].

146.   Sugama, Y., and A. B. Malik. Thrombin receptor 14-amino acid peptide mediates endothelial hyperadhesivity and neutrophil adhesion by P-selectin-dependent mechanism. Circ. Res. 71: 1015-1019, 1992[Abstract].

147.   Suidan, H. S., J. Bouvier, E. Schaerer, S. R. Stone, D. Monard, and J. Tschopp. Granzyme A released upon stimulation of cytotoxic T lymphocytes activates the thrombin receptor on neuronal cells and astrocytes. Proc. Natl. Acad. Sci. USA 91: 8112-8116, 1994[Abstract].

148.   Suidan, H. S., K. J. Clemetson, M. Brown-Luedi, S. P. Niclou, J. M. Clemetson, J. Tschopp, and D. Monard. The serine protease granzyme A does not induce platelet aggregation but inhibits responses triggered by thrombin. Biochem. J. 315: 939-945, 1996[Medline].

149.   Tay-Uyboco, J., M. C. Poon, S. Ahmad, and M. D. Hollenberg. Contractile actions of thrombin receptor-derived polypeptides in human umbilical and placental vasculature: evidence for distinct receptor systems. Br. J. Pharmacol. 115: 569-578, 1995[Abstract].

150.   Tesfamariam, B. Thrombin receptor-mediated vascular relaxation differentiated by a receptor antagonist and desensitization. Am. J. Physiol. 267 (Heart Circ. Physiol. 36): H1962-H1967, 1994[Abstract/Free Full Text].

151.   Tsuga, H., K. Kameyama, T. Haga, H. Kurose, and T. Nagao. Sequestration of muscarinic acetylcholine receptor m2 subtypes. Facilitation by G protein-coupled receptor kinase (GRK2) and attenuation by a dominant-negative mutant of GRK2. J. Biol. Chem. 269: 32522-32527, 1994[Abstract/Free Full Text].

152.   Van Biesen, T., L. M. Luttrell, B. E. Hawes, and R. J. Lefkowitz. Mitogenic signaling via G protein-coupled receptors. Endocr. Rev. 17: 698-714, 1996[Medline].

153.   Van Corven, E. J., P. L. Hordijk, R. H. Medema, J. L. Bos, and W. H. Moolenaar. Pertussis toxin-sensitive activation of p21ras by G protein-coupled receptor agonists in fibroblasts. Proc. Natl. Acad. Sci. USA 90: 1257-1261, 1993[Abstract].

154.   Vaughan, P. J., C. J. Pike, C. W. Cotman, and D. D. Cunningham. Thrombin receptor activation protects neurons and astrocytes from cell death produced by environmental insults. J. Neurosci. 15: 5389-5401, 1995[Abstract].

155.   Vaughan, P. J., J. Su, C. W. Cotman, and D. D. Cunningham. Protease nexin-1, a potent thrombin inhibitor, is reduced around cerebral blood vessels in Alzheimer's disease. Brain Res. 668: 160-170, 1994[Medline].

156.   Verrall, S., M. Ishii, M. Chen, L. Wang, T. Tram, and S. R. Coughlin. 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, 1997[Abstract/Free Full Text].

157.   Vouret-Craviari, V., P. Auberger, J. Pouyssegur, and E. Van Obberghen-Schilling. Distinct mechanisms regulate 5-HT2 and thrombin receptor desensitization. J. Biol. Chem. 270: 4813-4821, 1995[Abstract/Free Full Text].

158.   Vouret-Craviari, V., D. Grall, J. C. Chambard, U. B. Rasmussen, J. Pouyssegur, and E. Van Obberghen-Schilling. Post-translational and activation-dependent modifications of the G protein-coupled thrombin receptor. J. Biol. Chem. 270: 8367-8372, 1995[Abstract/Free Full Text].

159.   Vouret-Craviari, V., E. Van Obberghen-Schilling, J. C. Scimeca, E. Van Obberghen, and J. Pouyssegur. Differential activation of p44mapk (ERK1) by alpha -thrombin and thrombin-receptor peptide agonist. Biochem. J. 289: 209-214, 1993[Medline].

160.   Vu, T. K., D. T. Hung, V. I. Wheaton, and S. R. Coughlin. Molecular cloning of a functional thrombin receptor reveals a novel proteolytic mechanism of receptor activation. Cell 64: 1057-1068, 1991[Medline].

161.   Vu, T. K., V. I. Wheaton, D. T. Hung, I. Charo, and S. R. Coughlin. Domains specifying thrombin-receptor interaction. Nature 353: 674-677, 1991[Medline].

162.   Weinstein, J. R., S. J. Gold, D. D. Cunningham, and C. M. Gall. Cellular localization of thrombin receptor mRNA in rat brain: expression by mesencephalic dopaminergic neurons and codistribution with prothrombin mRNA. J. Neurosci. 15: 2906-2919, 1995[Abstract].

163.   Winitz, S., S. K. Gupta, N. X. Qian, L. E. Heasley, R. A. Nemenoff, and G. L. Johnson. Expression of a mutant Gi2alpha subunit inhibits ATP and thrombin stimulation of cytoplasmic phospholipase A2-mediated arachidonic acid release independent of Ca2+ and mitogen-activated protein kinase regulation. J. Biol. Chem. 269: 1889-1895, 1994[Abstract/Free Full Text].

164.   Yalkinoglu, A. O., P. Spreyer, M. Bechem, H. Apeler, and S. Wohlfeil. Induction of c-fos expression in rat vascular smooth muscle reporter cells by selective activation of the thrombin receptor. J. Recept. Signal Transduct. Res. 15: 117-130, 1995[Medline].

165.   Yu, Z., S. Ahmad, J. L. Schwartz, D. Banville, and S. H. Shen. Protein-tyrosine phosphatase SHP2 is positively linked to proteinase-activated receptor 2-mediated mitogenic pathway. J. Biol. Chem. 272: 7519-7524, 1997[Abstract/Free Full Text].

166.   Zacharias, U., Y. Xu, J. Hagege, J. D. Sraer, L. F. Brass, and E. Rondeau. Thrombin, phorbol ester, and cAMP regulate thrombin receptor protein and mRNA expression by different pathways. J. Biol. Chem. 270: 545-550, 1995[Abstract/Free Full Text].

167.   Zhang, J., W. G. King, S. Dillon, A. Hall, L. Feig, and S. E. Rittenhouse. Activation of platelet phosphatidylinositide 3-kinase requires the small GTP-binding protein Rho. J. Biol. Chem. 268: 22251-22254, 1993[Abstract/Free Full Text].


Am J Physiol Cell Physiol 274(6):C1429-C1452
0002-9513/98 $5.00 Copyright © 1998 the American Physiological Society