Download Print

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Subventricular zone wikipedia , lookup

Optogenetics wikipedia , lookup

Neuromuscular junction wikipedia , lookup

NMDA receptor wikipedia , lookup

Synaptogenesis wikipedia , lookup

Stimulus (physiology) wikipedia , lookup

Molecular neuroscience wikipedia , lookup

Channelrhodopsin wikipedia , lookup

Endocannabinoid system wikipedia , lookup

Clinical neurochemistry wikipedia , lookup

Neuropsychopharmacology wikipedia , lookup

Signal transduction wikipedia , lookup

Transcript
Physiol Rev
84: 579 – 621, 2004; 10.1152/physrev.00028.2003.
Protease-Activated Receptors:
Contribution to Physiology and Disease
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
Departments of Surgery and Physiology, University of California, San Francisco, California
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
I. Introduction
II. Protease-Activated Receptors: Mechanisms of Activation by Cell-Surface Proteolysis
A. Discovery and structure
B. Proteases cleave PARs to expose a tethered ligand domain
C. Protease binding to PARs facilitates cleavage and activation
D. There are functional interactions between PARs
E. Protease binding to other membrane proteins can facilitate PAR cleavage and activation
F. Multiple proteases activate PARs
III. Protease-Activated Receptors: Transduction of the Signal
A. Tethered ligands interact with extracellular domains to initiate signal transduction
B. Structure-activity relations of tethered ligand domains
C. PARs activate multiple signaling cascades
IV. Protease-Activated Receptors: Termination of the Signal
A. Cell-surface proteolysis can disable PARs
B. PAR signaling is attenuated by receptor desensitization
C. PARs are downregulated by intracellular proteases
D. Sustained signaling by proteases requires mobilization and synthesis of new receptors
V. Protease-Activated Receptors: Contribution to Physiological and Pathophysiological
Control Mechanisms
A. Protease signaling in the circulatory and cardiovascular systems
B. Protease signaling to the nervous system
C. Protease signaling in the gastrointestinal system
D. Protease signaling in the airway
E. Protease signaling in the skin
VI. Protease-Activated Receptors: Mechanisms of Disease and Targets for Therapy
A. Cardiovascular disease
B. Inflammatory disease
C. Cancer
VII. Conclusions and Future Perspectives
580
581
581
582
583
584
585
587
589
589
590
590
594
594
594
596
597
598
599
601
604
607
609
610
610
611
611
612
Ossovskaya, Valeria S., and Nigel W. Bunnett. Protease-Activated Receptors: Contribution to Physiology
and Disease. Physiol Rev 84: 579 – 621, 2004; 10.1152/physrev.00028.2003.—Proteases acting at the surface of
cells generate and destroy receptor agonists and activate and inactivate receptors, thereby making a vitally
important contribution to signal transduction. Certain serine proteases that derive from the circulation (e.g.,
coagulation factors), inflammatory cells (e.g., mast cell and neutrophil proteases), and from multiple other
sources (e.g., epithelial cells, neurons, bacteria, fungi) can cleave protease-activated receptors (PARs), a family
of four G protein-coupled receptors. Cleavage within the extracellular amino terminus exposes a tethered ligand
domain, which binds to and activates the receptors to initiate multiple signaling cascades. Despite this
irreversible mechanism of activation, signaling by PARs is efficiently terminated by receptor desensitization
(receptor phosphorylation and uncoupling from G proteins) and downregulation (receptor degradation by
cell-surface and lysosomal proteases). Protease signaling in tissues depends on the generation and release of
proteases, availability of cofactors, presence of protease inhibitors, and activation and inactivation of PARs.
Many proteases that activate PARs are produced during tissue damage, and PARs make important contributions
to tissue responses to injury, including hemostasis, repair, cell survival, inflammation, and pain. Drugs that
mimic or interfere with these processes are attractive therapies: selective agonists of PARs may facilitate
www.prv.org
0031-9333/04 $15.00 Copyright © 2004 the American Physiological Society
579
580
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
healing, repair, and protection, whereas protease inhibitors and PAR antagonists can impede exacerbated
inflammation and pain. Major future challenges will be to understand the role of proteases and PARs in
physiological control mechanisms and human diseases and to develop selective agonists and antagonists that
can be used to probe function and treat disease.
I. INTRODUCTION
FIG. 1. Mechanisms by which cell-surface
proteolysis regulates signal transduction. Proteases can regulate signaling by cleaving ligands
(A–C) or receptors (D and E). A: cell-surface
proteases can induce shedding of membranebound signaling molecules; e.g., tumor necrosis
factor-␣ (TNF-␣)-converting enzyme or TACE, a
member of the ADAM (a disintegrin and metalloproteinase) family of cell-surface proteases,
liberates soluble TNF-␣ from the cell surface
and thus releases a soluble cytokine. B: cellsurface peptidases can generate biologically active peptides; e.g., angiotensin converting enzyme or ACE converts the decapeptide angiotensin I (ANG I) to the octapeptide ANG II, the
principal active form. C: cell-surface peptidases
can degrade and inactivate neuropeptides; e.g.,
neutral endopeptidase (NEP) cleaves substance
P (SP) at multiple sites to form inactive fragments. D: soluble proteases can cleave proteaseactivated receptors (PARs) to expose a tethered
ligand that binds and activates the cleaved receptor; e.g., the coagulation factor thrombin
cleaves PAR1 to activate the receptor. E: soluble
proteases can cleave PARs to remove the tethered ligand, generating a disabled receptor; e.g.,
cathepsin G from neutrophils cleaves PAR1 to
remove the tethered ligand and thereby prevent
activation by thrombin.
Physiol Rev • VOL
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
Proteolytic enzymes are estimated to comprise 2% of
the human genome (266). It is not surprising, therefore,
that proteases participate in a diverse array of biological
processes, ranging from degradation of dietary proteins in
the lumen of the gastrointestinal tract to the control of the
cell cycle. This review concerns the importance of proteolytic events at the cell surface in signal transduction,
physiological control, and disease.
Proteases that are anchored to the plasma membrane
or that are soluble in the extracellular fluid can cleave
ligands or receptors at the surface of cells to either initiate or terminate signal transduction (Fig. 1). Thus cellsurface proteases can release or generate active ligands,
or degrade and inactivate receptor agonists. For example,
tumor necrosis factor (TNF)-␣-converting enzyme or
TACE cleaves the precursor of TNF-␣ at the plasma membrane, thereby releasing a soluble form of this proinflammatory cytokine. In a similar manner, angiotensin converting enzyme (ACE), which is also an integral membrane protein, converts angiotensin I to angiotensin II in
the extracellular fluid to generate the principal active
form of this hormone. In contrast, neutral endopeptidase
degrades and inactivates the neuropeptide substance P
(SP) in the vicinity of its receptors and thus terminates
the biological effects of SP. Certain soluble and membrane-bound proteases cleave G protein-coupled receptors (GPCRs) at the cell surface to activate or inactivate
receptors. For example, the coagulation factor thrombin
cleaves protease-activated receptor 1 (PAR1) on platelets,
which activates the receptor to induce platelet aggregation and hemostasis. Conversely, cathepsin G from neutrophils cleaves PAR1 at a different site from thrombin to
581
PROTEASE-ACTIVATED RECEPTORS
II. PROTEASE-ACTIVATED RECEPTORS:
MECHANISMS OF ACTIVATION BY
CELL-SURFACE PROTEOLYSIS
A. Discovery and Structure
1. PAR1
PAR1, formerly known as the thrombin receptor, was
cloned by two laboratories by the strategy of expressing
RNA from thrombin-responsive cells of humans and hamsters in oocytes from Xenopus (233, 305). Clones were
identified that encoded a protein of 425 residues with 7
hydrophobic domains of a typical GPCR. The deduced
sequence of human PAR1 contained an amino-terminal
signal sequence, and extracellular amino-terminal domain
of 75 residues, and a potential cleavage site for thrombin
within the amino-tail (LDPR412S42FLLRN, where 2 denotes cleavage) (Fig. 2).
2. PAR2
PAR1 remained the only member of the family until
the serendipitous discovery of PAR2. PAR2 was identified
by screening a mouse genomic library using degenerate
primers to the second and sixth transmembrane domains
of the bovine neurokinin 2 receptor (214, 215). A clone
was found that encoded a protein of 395 residues with the
1. A summary of PAR activating proteases, disabling proteases, activating peptides, localization, and
phenotypes of PAR-deficient mice
TABLE
PAR1
Activating
proteases
Thrombin
FXa
APC
Granzyme A
Gingipains-R
Trypsin
Inactivating
proteases
Cathepsin G
Plasmin
Elastase
Proteinase-3
Trypsin
LDPR41 2 S42FLLRN
SFLLRN
TFLLRN
Platelets (human),
endothelium, epithelium,
fibroblasts, myocytes,
neurons, astrocytes
Partial embryonic lethality,
vascular matrix
deposition after injury
Cleavage site
AP
Localization
Phenotype of
knockout
PAR2
PAR3
PAR4
Trypsin
Tryptase
FVIIa
FXa
MT-SP1
Proteinase-3
Acrosien
Der P3 D9
Gingipains-R
Elastase
Cathepsin G
Thrombin
SKGR34 2 S35LIGKV
SLIGKV
LPIK38 2 T39FRGAP
None
Epithelium, endothelium,
fibroblasts, myocytes,
neurons, astrocytes
Platelets (mouse),
endothelium,
myocytes, astrocytes
PAPR47 2 G48YPGQV
GYPGQV
AYPGKF
Platelets (human),
endothelium,
myocytes, astrocytes
Impaired leukocyte
migration; impaired
allergic inflammation of
airway, joints, kidney
Protection against
thrombus formation/
pulmonary embolism
Protection against
thrombus formation/
pulmonary embolism
Cathepsin G
References are provided in the text. PAR, protease-activated receptor; AP, activating peptide.
Physiol Rev • VOL
84 • APRIL 2004 •
Thrombin
Trypsin
Cathepsin G
Gingipains-R
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
generate a disabled receptor that cannot respond to
thrombin, which could impede blood clotting. These proteolytic events are critically important for normal physiological control: the conversion of angiotensin I to angiotensin II is central to the reflex regulation of blood pressure and volume, and thrombin-induced aggregation of
platelets is vital for normal hemostasis. However, these
processes are also of great interest in understanding
mechanisms of disease and for the development of effective therapies. Thus inhibitors of ACE are widely used to
treat hypertension and congestive heart failure, and antagonists of PARs are being developed to treat thrombotic
and inflammatory diseases.
The focus of this review is on the PAR family
GPCRs. Four PARs have been identified by molecular
cloning (Table 1). A wide range of proteases cleave and
activate PARs, including proteases from the coagulation cascade, inflammatory cells, and the digestive
tract. Receptor activation initiates an array of signaling
events in many cell types with diverse consequences,
ranging from hemostasis to pain transmission. We review the mechanisms by which cell-surface proteolysis
activates PARs to initiate signal transduction, discuss
the mechanisms that terminate signaling by PARs, review the importance of PARs in physiological control in
major systems, and speculate on the contribution of
PARs to disease. There are several comprehensive reviews of PARs (68, 108, 182, 217).
582
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
typical characteristics of a GPCR and with ⬃30% amino
acid identity to human PAR1. The extracellular amino
terminus of 46 residues contained a putative trypsin cleavage site SKGR342S35LIGKV (Fig. 2).
3. PAR3
The discovery of PAR2 provided the impetus for attempts to identify other receptors of this type. Indeed, the
existence of additional receptors for thrombin was suggested by the observation that platelets derived from
PAR1-deficient mice still responded to thrombin, whereas
fibroblasts were unresponsive (64). PAR3 was subsequently cloned using degenerate primers to conserved
domains of PAR1 and PAR2 to screen RNA from rat platelets (135). The human and murine forms of PAR3 were
subsequently cloned, and human PAR3 was found to share
⬃28% sequence homology to human PAR1 and PAR2. Like
PAR1 and PAR2, PAR3 is a typical GPCR with a thrombin
cleavage site within the extracellular amino terminus at
LPIK382T39FRGAP (Fig. 2).
groups subsequently used these sequences to clone a new
GPCR, PAR4 (143, 317). Human PAR4 is a 385-amino acid
protein, with a potential cleavage site for thrombin and
trypsin in the extracellular amino-terminal domain:
PAPR472G48YPGQV (Fig. 2). PAR4 is ⬃33% homologous to
the other human PARs, but with some distinct differences in
the amino- and carboxy-terminal domains.
B. Proteases Cleave PARs to Expose a Tethered
Ligand Domain
The general mechanism by which proteases cleave and
activate PARs is the same: proteases cleave at specific sites
within the extracellular amino terminus of the receptors;
this cleavage exposes a new amino terminus that serves as
a tethered ligand domain, which binds to conserved regions
in the second extracellular loop of the cleaved receptor,
resulting in the initiation of signal transduction (Fig. 2).
There is no known function of the amino-terminal fragment
of the receptor that is removed by proteolysis.
4. PAR4
1. PAR1
Two laboratories identified PAR-like sequences by
searching expressed sequence tag (EST) libraries, and both
The mechanism by which thrombin activates PAR1
has been investigated in detail. Thrombin cleaves PAR1 at
Physiol Rev • VOL
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
FIG. 2. Structural and functional domains of PARs. The figure shows alignment of domains of human PAR1, PAR2,
PAR3, and PAR4. A and B: mechanism of
cleavage and interaction of the tethered
ligand with extracellular binding domains. C: functionally important domains
in the amino terminus, second extracellular loop, and carboxy terminus. Conserved residues in loop II are in bold.
[Adapted from Derian et al. (89) and Macfarlane et al. (182).]
PROTEASE-ACTIVATED RECEPTORS
2. PAR2
Similar observations indicate that trypsin cleaves
PAR2 at R342S35LIGKV to reveal the amino-terminal tethered ligand SLIGKV in humans (213–215). Mutation of the
trypsin site prevents trypsin cleavage and activation of
PAR2. Synthetic peptides corresponding to the tethered
ligand domain (SLIGKV) activate PAR2 without the need
for receptor cleavage. Exposure of cells to trypsin results
in loss of immunoreactivity to an antibody against an
amino-terminal epitope, which indicates that trypsin
cleaves intact PAR2 at the cell surface (23).
3. PAR3
Thrombin cleaves PAR3 at K382T39FRGAP, and mutation of the cleavage site to one that would be resistant
to thrombin prevents activation (135). Cleavage by thrombin exposes a new amino terminus (TFRGAP) that may
interact with the receptor as a tethered ligand. However,
in marked contrast to PAR1, PAR2, and PAR4, synthetic
peptides corresponding to this putative tethered ligand do
not activate PAR3. The reason for this discrepancy is
unknown, although differences in affinity, steric hindrances, and the possibility that cleavage releases conformation of the receptor constrained by the uncleaved
amino-terminal region could explain these unexpected
results. Another unexpected and unexplained observation
is that mouse PAR3 is unable to signal when expressed
alone, without other PARs (143).
Physiol Rev • VOL
4. PAR4
Thrombin and trypsin cleave PAR4 at R472G48YPGQV,
and peptides corresponding to the tethered ligand domain
GYPGQV can directly activate PAR4 (143, 317). Mutation of
the cleavage site prevents activation by thrombin and trypsin, but not by the synthetic peptide, which confirms the
importance of proteolytic cleavage for receptor activation.
C. Protease Binding to PARs Facilitates Cleavage
and Activation
Thrombin can activate PAR1 and PAR3 by a two-step
mechanism: first the protease binds and then it cleaves
the receptor (Fig. 3A). The process of binding and activation has been most thoroughly studied for thrombin and
PAR1 (306). The extracellular amino terminus of human
PAR1 contains a sequence of charged residues
(D51KYEPF56) that is distal to the thrombin cleavage site.
This charged domain binds to an anion binding site on
thrombin, thereby temporarily concentrating the protease
at the surface of the receptor. This negatively charged
region of PAR1 resembles a domain of the leech anticoagulant hirudin, which inhibits thrombin by binding its
anion site. The importance of the hirudin-like domain is
emphasized by the finding that its deletion markedly diminishes the capacity of thrombin to activate PAR1,
whereas substitution of this region with the corresponding domain of hirudin allows a full recovery of activity.
Moreover, ␥-thrombin, which lacks an anion site, is 100fold less potent than ␣-thrombin, which has the site, in
cleaving PAR1 at the activation site (25). Furthermore,
platelets respond to low concentrations of ␣-thrombin but
not ␥-thrombin, because the latter cannot bind to the
PAR1 (260). Protease binding thereby increases the efficiency of activation, probably by concentrating the protease on the surface of the receptor or by altering the
conformation of the receptor to facilitate cleavage. PAR3
also contains a hirudin-like site (FEEFP) that is distal to
the thrombin cleavage site, which interacts with thrombin
(135). Thus ␥-thrombin is 100-fold less potent than
␣-thrombin for activating PAR3, and alanine substitutions
within the hirudin site of PAR3 attenuate activation of
PAR3 by thrombin. PAR4, in contrast to the other thrombin receptors, lacks a hirudin-like binding site for thrombin (143, 317). It is for this reason that both ␣-thrombin
and ␥-thrombin can activate PAR4 with similar potency
(317). The lack of a thrombin binding site also accounts
for the observation that ␣-thrombin activates PAR4 with a
potency that is 50-fold less than for activation of PAR1.
Thus PAR4 is a low-affinity receptor for thrombin,
whereas PAR1 and PAR3 are high-affinity thrombin receptors. Protease binding sites have not been identified for
other PARs.
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
R412S42FLLRN to expose the tethered ligand SFLLRN,
which binds and activates the cleaved receptor, resulting
in signal transduction. Several observations support this
mechanism of activation. Mutation of the cleavage site
prevents thrombin cleavage and signaling, indicating the
importance of this site for activation of PAR1 (305). The
general importance of proteolytic activation is indicated
by the finding that replacement of thrombin site with an
enterokinase site generates a receptor responsive to enterokinase (306). A synthetic peptide that mimics the
tethered ligand domain (S42FLLRNPNDKYEPF) directly
activates intact PAR1, without the requirement for hydrolysis by thrombin (305), and peptides as short as six
residues (S42FLLRN) are also fully active (253) (see sect.
IIIB). Such synthetic agonists, referred to as activating
peptides (AP), are useful tools for investigating PAR functions. Direct physical proof that thrombin cleaves PAR1 is
provided by the findings that exposure of platelets to
thrombin reduces binding an antibody directed against
the cleavage site of PAR1 but does not alter binding of an
antibody directed to a domain distal to the cleavage site
(212). Moreover, exposure to thrombin increases the electrophoretic mobility of epitope-tagged PAR1, indicating a
reduction of molecular weight by proteolytic cleavage
(304).
583
584
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
D. There Are Functional Interactions
Between PARs
A common theme of signaling by GPCRs is that there
are frequently several different receptors for a single ligand and often several ligands for one receptor. Thus
thrombin can activate PAR1, PAR3, and PAR4, with different potencies, and trypsin, tryptase, and certain coagulation factors can activate PAR2. This complexity becomes
particularly interesting because PARs are frequently coexpressed, and interactions between receptors in the
same cell can have important functional consequences.
1. Dual receptor systems
Human platelets express two thrombin receptors:
PAR1 and PAR4. There are marked differences in the
mechanisms of activation and inactivation of these receptors that have important consequences for thrombin signaling to platelets. PAR1, by virtue of a hirudin-like site, is
able to bind thrombin and thus responds to low concentrations of enzyme. PAR4 lacks the hirudin site and can
Physiol Rev • VOL
respond only to high thrombin concentrations (142).
However, whereas PAR1 responses are rapidly shut-off,
probably as a result of phosphorylation of residues in the
carboxy terminus and uncoupling from G proteins (see
sect. IV), PAR4 responses are sustained and thus desensitize slowly (263). The coexpression of two receptors with
different potencies and kinetics of desensitization may
have important functional consequences. Thus PAR1 mediates rapid and transient increases in intracellular Ca2⫹
concentration ([Ca2⫹]i) in human platelets to low concentrations of thrombin, whereas PAR4 mediates delayed and
sustained increases in [Ca2⫹]i to higher thrombin concentrations (70). This prolonged signal is important for the
late phases of platelet aggregation. Similar dual receptor
systems probably exist on other cells with important functional consequences.
2. PAR3 as a cofactor for PAR4
In some instances, protease binding to one receptor
can facilitate cleavage of another receptor that is expressed on the same cell. This appears to be the situation
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
FIG.
3. Mechanisms of activation
and intermolecular cooperation of PARs.
A: thrombin activates PAR1 and PAR3 in a
two-step process. First, thrombin binds
to a hirudin-like domain; second, thrombin cleaves to expose the tethered ligand,
which binds and activates the cleaved
receptor. B: PAR3 is a cofactor for PAR4
in murine platelets. Thrombin binds to
the hirudin site of PAR3, but PAR3 does
not signal in mouse platelets. The PAR3bound thrombin then cleaves and activates PAR4. C: intermolecular signaling
of PAR2 and PAR1 in endothelial cells.
The exposed tethered ligand domain of
PAR1 can bind and activate PAR2, resulting in intermolecular signaling. [Adapted
from Coughlin (68).]
PROTEASE-ACTIVATED RECEPTORS
3. Intermolecular PAR signaling
GPCRs can form homodimers and heterodimers with
important consequences for signal transduction. Although
the principal mechanism of PAR activation is intramolecular (i.e., the unmasked tethered ligand binds to the
cleaved receptor), there are several examples of intermolecular interactions between different PAR molecules. Intermolecular signaling, by which a cleaved receptor can
activate an uncleaved receptor, was first demonstrated
for PAR1 (43). The approach was to express the amino
terminus of PAR1 (including the tethered ligand domain)
attached to the transmembrane domain of CD8 (to anchor
the fragment at the cell surface) with truncated PAR1
lacking the amino terminus, either alone or together.
When expressed alone, there was no response to thrombin. However, when coexpressed, cells were able to respond to thrombin, suggesting that thrombin cleaves the
amino terminus of PAR1 to reveal the tethered ligand
domain, which then interacts with the truncated receptor,
which transduces the signal. Thus at least in a reconstituted system, it is possible for the tethered ligand of
cleaved PAR1 to activate an uncleaved receptor. There is
Physiol Rev • VOL
also evidence of intermolecular signaling between different PARs (Fig. 3C). Peptides corresponding to the tethered ligand of PAR1 (SFLLRN) can also activate PAR2, but
not vice versa (21). To examine intermolecular signaling,
a signaling defective mutant of PAR1 and wild-type PAR2
were expressed alone or together (218). When expressed
alone, neither receptor responded to thrombin, but when
coexpressed, there was a robust thrombin response.
Moreover, in endothelial cells that naturally express PAR1
and PAR2, a PAR1 antagonist blocked only 75% of the
response to thrombin, whereas desensitization of PAR2
blocked the remaining 25% of the response. These results
suggest that the tethered ligand domain of PAR1 can
interact with uncleaved PAR2 to transduce the signal. This
novel form of intermolecular signaling between different
PARs clearly requires the close association of receptors at
the cell surface, which could be influenced by levels of
expression or by anchoring proteins that may affect mobility of receptors in the membrane.
E. Protease Binding to Other Membrane Proteins
Can Facilitate PAR Cleavage and Activation
As discussed, thrombin binding to a PAR can facilitate cleavage and activation of that receptor or a different
PAR. In addition, binding to nonreceptor proteins can
concentrate proteases at the cell surface and thereby
facilitate PAR activation. For example, glycoprotein I␤␣
is a cell-surface platelet protein with a high-affinity binding site for ␣-thrombin. Disruption of this interaction
impedes the capacity of ␣-thrombin to cleave PAR1 and
cause aggregation of platelets, suggesting that glycoprotein I␤␣ is a cofactor for PAR1 that promotes PAR1mediated platelet aggregation (84). Thrombin may also
interact with other proteins on platelets (118).
This theme of anchoring proteins serving as cofactors for proteolytic activation of PARs is particularly important for signaling by coagulation factors (F), notably
FVIIa and FXa that act upstream of thrombin (reviewed in
Refs. 68, 240). Coagulation proceeds by two processes: an
extrinsic and an intrinsic pathway. The extrinsic pathway
is initiated by tissue damage and requires tissue factor
(TF) or thromboplastin, an integral membrane protein
that is normally expressed by extravascular cells (e.g.,
myocytes, keratinocytes), and which is expressed by endothelial cells and monocytes during inflammation. During coagulation, TF binds FVIIa, and the TF-FVIIa complex interacts with the zymogen FX to form active FXa.
FXa then interacts with FVa, a cofactor for FXa-mediated
conversion of prothrombin to thrombin. Thrombin plays a
critical role in hemostasis by converting fibrinogen to
fibrin and by cleaving PARs on platelets to induce aggregation. The intrinsic pathway is triggered by exposure of
blood to collagen underlying damaged endothelium, and
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
for PAR3 and PAR4, the thrombin receptors on murine
platelets (mouse platelets lack PAR1) (199) (Fig. 3B).
Observations in mice deficient in PAR3 indicate that PAR3
is required for the response of platelets to low but not to
high concentrations of thrombin, which is attributable to
PAR4. These findings could be explained by the fact that
PAR3 contains a thrombin binding site, and can thus
respond to low concentrations of thrombin (135),
whereas PAR4 lacks a binding site and thus mediates
responses to high thrombin concentrations (143, 317).
However, the situation is complicated by the finding that
murine PAR3 expressed in COS cells does not signal, for
reasons that are not fully understood. This finding suggests that, in murine platelets, PAR3 somehow facilitates
the activation of PAR4 by low concentrations of thrombin.
In support of this hypothesis, the potency of thrombin
signaling to PAR4 in COS cells is increased 6- to 15-fold by
coexpression of PAR3, whereas responses to PAR4 AP are
unaffected (199). Moreover, the expression of a construct
of the amino terminus of PAR3 (including the hirudin-like
site) attached to the transmembrane domain of CD8 (to
anchor the construct at the cell surface) with PAR4 also
facilitates thrombin signaling. Together, these findings
suggest that PAR3 is a cofactor for PAR4 in murine platelets: the hirudin-like site of PAR3 binds and concentrates
thrombin at the cell surface, and thereby promotes thrombin cleavage of PAR4. At high concentrations, thrombin
can directly cleave and activate PAR4, even though it
lacks a thrombin binding domain. This novel mechanism
of cooperative interaction between PAR3 and PAR4 adds
to the complexity of interactions between GPCRs.
585
586
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
the presence of FX. These results suggest the local formation of FXa at the plasma membrane results in PAR2
activation (34). The receptor that mediates FXa signaling
depends on the cell in question. Observations made using
cells prepared from PAR-deficient animals indicate that
FXa signaling to endothelial cells is mostly mediated by
PAR2 but also by PAR1, whereas PAR1 solely mediates
FXa signaling to fibroblasts (35). These mechanisms of
signaling by FVIIa and FXa may be of particular importance during septic shock when upstream coagulation
proteases make a substantial contribution to the inflammatory response and when the expression of TF is upregulated on several cell types.
The role of anchoring proteins in PAR signaling has
also been extended to the anticoagulant pathway (Fig.
4B). Low levels of thrombin are normally present in the
circulation, and when vessels remain intact these low
levels of thrombin contribute to the anticoagulant protein
C (APC) pathway. The endothelial protein C receptor
(EPCR) binds protein C to the surface of endothelial cells.
Thrombomodulin also binds thrombin to the endothelial
cell surface, where thrombin converts protein C to APC.
When APC is released from the EPCR, it acts as a circu-
FIG. 4. Formation of coagulation and
anticoagulation complexes promotes activation of PARs. A: tissue factor (TF)
anchors FVIIa to the cell surface and
thereby facilitates activation of PAR1 and
PAR2. A ternary complex of TF-FVIIaFXa efficiently cleaves and activates
PAR1 and PAR2. B: thrombomodulin
(TM) concentrates low concentrations of
thrombin to the surface of endothelial
cells. Endothelial protein C receptor
(EPCR) binds protein C (PC), and thrombin converts PC to activated PC (APC).
APC, still bound to EPCR, then activates
PAR1.
Physiol Rev • VOL
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
begins with conversion of FXII to FXIIa, which is catalyzed by kallikrein and kininogen. This conversion then
initiates a cascade of events involving activation of FXI,
FX, and converging with the extrinsic pathway and the
FXa-dependent conversion of prothrombin to thrombin.
In addition to zymogen cleavage, FVIIa and FXa can
signal to PARs. FVIIa signals to cells when allosterically
associated with TF for full activity, in part through PAR1
and PAR2 (34) (Fig. 4A). Thus, although FVIIa, even at
high concentrations, does not signal to Xenopus oocytes
expressing PAR1 or PAR2, FVIIa robustly signals to oocytes expressing TF together with PAR1 or PAR2. Similarly, soluble FXa weakly activates PAR1, PAR2, or PAR4,
but when FXa is associated with a complex comprising
TF-FVIIa-FXa, it potently activates PAR1 and PAR2 (34,
239). Moreover, although the TF-FVIIa complex can activate PAR2, it does so with far lower efficiency than the
TF-FVIIa-FXa complex, in which locally generated FXa
induces signaling. Thus FVIIa can signal to fibroblasts
expressing PAR2 and TF with a potency of ⬃4 nM, but this
concentration exceeds the estimated plasma concentration of FVIIa (50 pM). However, the potency of FVIIa is
reduced by almost three orders of magnitude (to 8 pM) in
PROTEASE-ACTIVATED RECEPTORS
F. Multiple Proteases Activate PARs
Although many proteases have been found to cleave
and activate PARs in cultured cells, the capacity of a
protease to signal in intact tissues depends on many
factors. First, proteases must be generated or released in
sufficient concentrations to activate PARs. Although the
catalytic properties of proteases would ensure that even
very low concentrations could eventually cleave all receptors on the surface of the cell, it is the rate of hydrolysis
of PARs that determines the magnitude of the resulting
cellular signal (137). Second, efficient hydrolysis and activation of PARs may require the presence of accessory
cofactors, for example, TF and EPCR in the case of
FVIIa-FXa and APC, respectively (34, 46, 238, 239). Third,
the capacity of a protease to signal will depend on the
availability of protease inhibitors that serve to dampen
the effects of many proteases in vivo. Thus, although
trypsin is a very potently activated PAR2 in cultured cells,
trypsin inhibitors are widely expressed and may well limit
the capacity of trypsin to signal in tissues. With these
caveats in mind, many proteases have been identified that
are capable, at least theoretically, of activating PARs.
Physiol Rev • VOL
1. Coagulation factors
Serine proteases of the coagulation cascade are perhaps the best characterized activators of PARs (68, 108,
240). As discussed in section IIE, proteases that mediate
coagulation and anticoagulation by cleaving zymogens or
active enzymes themselves can also signal to several cell
types by cleaving and activating PARs. Thus thrombin
activates PAR1, PAR3, or PAR4 at the surface of platelets,
resulting in aggregation, which contributes to hemostasis
(68). The TF-FVIIa-FXa complex signals by cleaving PAR1
and PAR2 on a variety of cell types, including endothelial
cells, which is of particular importance in inflammation
(34, 239). Thrombin can also exert anti-inflammatory effects that depend on the local generation of APC and
consequent activation of PAR1 on endothelial cells (46,
238).
2. Pancreatic and extrapancreatic trypsins
Trypsins potently activate PAR2 and PAR4. There are
at least three distinct trypsin genes in humans: trypsin I,
II, and mesotrypsin; trypsin IV is a splice variant of mesotrypsin. The potential of trypsins to signal to cells by
cleaving PAR2 or PAR4 depends on the release of the
zymogen trypsinogen, the presence of enteropeptidase,
which activates trypsinogen, and the existence of the
large array of endogenous trypsin inhibitors. During feeding, trypsinogens I and II are secreted from the pancreas
into the lumen of the small intestine, where they are
activated by enteropeptidase. In feeding rats, luminal
trypsin attains concentrations (1 ␮M) that are more than
capable of strongly activating PAR2 at the apical surface
of enterocytes (EC50 ⬃5 nM) (167). Pancreatic trypsinogens are also prematurely activated in the inflamed pancreas where they are released into the interstitial fluid and
vasculature and could activate PAR2 in pancreatic acini,
duct cells, or nerves (205). However, trypsins are widely
distributed enzymes that are expressed by many extrapancreatic cells, including endothelial cells (169), epithelial cells, the nervous system (168), and in tumors (165,
166). However, despite this widespread distribution, almost nothing is known about the control of secretion or
activation of extrapancreatic trypsins or their potential
functions as PAR activators. Trypsin II isolated from conditioned medium from colon cancer cell lines can cleave
and activate PAR2; since these cells also express PAR2 it
is theoretically possible that trypsin II could regulate cells
in an autocrine manner (6, 94). A tryptic-like serine protease purified from rat brain, designated P22, degrades
matrix and can signal to cells by activating PAR2 (252).
P22 could be of particular interest in brain injury, where
there appears to be enhanced secretion. Trypsin IV is also
a potential PAR activator. Trypsinogen IV is invariably
coexpressed in epithelial cell lines, endothelial cells, and
human colonic mucosa with PAR2, and trypsin IV cleaves
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
lating anticoagulant protein by degrading FVa and FVIIa.
However, when retained at the cell surface, the EPCRAPC complex promotes activation of PAR1 (238). Thus
APC can signal to fibroblasts only when they coexpress
EPCR and PAR1, and APC signaling to endothelial cells is
blocked by an active site blocked form of APC, which
competes for EPCR, and by antibodies that block PAR1
cleavage. Gene expression analysis of endothelial cells
indicates that this mechanism of APC-induced activation
of PAR1 mediates in large part the anti-inflammatory effects of APC. Thus PAR1 mediates the effects of low
concentrations of APC on gene expression in endothelial
cells. In addition to its anticoagulant activity, APC is an
anti-inflammatory agent that reduces organ damage in
animal models (111) and reduces mortality of patients
with sepsis (20). PAR1 can contribute to the anti-inflammatory actions of APC. Thus APC prevents hypoxic-induced apoptosis of brain endothelial cells through transcriptionally dependent inhibition of tumor suppressor
gene p53, regulation of Bax/Bcl2 proteins, and reduction
of caspase-3 signaling (46). Blockade of PAR1 with an
antibody against the activation site prevents APC-mediated cytoprotection in cultured endothelial cells, indicating a crucial role for PAR1 in this protection. In a model of
focal ischemic stroke in mice, APC markedly reduces
brain infarction volumes. This effect is attenuated in
EPCR-deficient mice, and the neuroprotective role of APC
is abolished by PAR1 blocking antibodies. Thus EPCR and
PAR1 play a major role in the protective effects of APC.
587
588
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
and activates PAR2 and PAR4 (N. W. Bunnett, unpublished
observations). Of particular interest, trypsin IV is resistant
to most proteinaceous trypsin inhibitors that effectively
inhibit trypsins I and II (146). Together, these results raise
the intriguing possibility that trypsin IV can signal for
prolonged periods in tissues by autocrine or paracrine
activation of PAR2 and PAR4.
3. Mast cell proteases
Physiol Rev • VOL
4. Leukocyte proteases
Proteases from leukocytes are released at sites of
inflammation and may serve as PAR activators under
these conditions. Neutrophils store a variety of proteases
(cathepsin G, elastase, proteinase 3) in azurophil granules, which may be released on activation. Cathepsin G is
released from activated neutrophils and causes aggregation of platelets. This effect of cathepsin G may be mediated by PAR4 (249). Thus cathepsin G increases [Ca2⫹]i in
PAR4-transfected fibroblasts, PAR4-expressing oocytes,
and human platelets, and a PAR4-blocking antibody inhibits activation of platelets by cathepsin G and prevents
Ca2⫹ signaling in platelets induced by activation of neutrophils. Thus cathepsin G may activate platelets and
other cells at sites of injury and inflammation by cleaving
PAR4. However, some of the effects of cathepsin G are not
PAR mediated, since cathepsin G also signals to cardiac
myocytes in a PAR-independent manner (245). Proteinase
3 is present in neutrophil secretory granules and at the
cell surface and is also expressed by other inflammatory
cells. Proteinase 3 cleaves peptide fragments of PAR2 at
the activation site, and proteinase 3-induced Ca2⫹ responses in oral epithelial cells are suppressed by desen-
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
There has been considerable interest in mast cell
tryptase as an activator of PAR2. Tryptase is the most
abundant protease of human mast cells; it comprises up
to 25% of the total cellular proteins and is expressed by
almost all subsets of human mast cells (38). However,
there are distinct interspecies differences in the expression of proteases in mast cells (reviewed in Ref. 190).
Thus, whereas tryptase appears to be abundant in the
guinea pig (188), it is a less prominent protease in mast
cells of rats and mice. Many of the proinflammatory and
mitogenic effects of tryptase are mimicked by PAR2 APs,
suggesting that tryptase exerts its effects through this
receptor. However, there is some unresolved controversy
about the effectiveness of tryptase as a PAR2 activator.
On one hand, human tryptase from a variety of sources
(purified from human lung, skin, and mast cell lines) can
cleave PAR2 to expose the tethered ligand domain and
signals to transfected cells as well as many cell types that
naturally express PAR2 at physiological levels (2, 18, 19,
66, 67, 193, 196, 235, 254, 256, 268, 270). These signals
appear to be mediated by tryptase, for they are abolished
by selective inhibitors, and are PAR2 dependent because
they are absent from nontransfected cells and are diminished by selective downregulation of PAR2. The general
consensus of these studies is that tryptase is a PAR2
activator but that it is considerably less potent than trypsin. However, most of the preparations used in these
studies likely contain several forms of tryptase. Human
mast cells express at least five distinct tryptase genes: ␣,
␤I, ␤II, ␤III, and transmembrane tryptase, and splice variants also exist. Thus the molecular form of tryptase responsible for these effects is unknown. Observations using recombinant tryptase are less clear. Although recombinant ␣ and ␤II tryptase have been reported to activate
PAR2 in BaF3 cells (193), another report found no activity
of ␤I tryptase as a PAR activator (130). The reason for this
discrepancy is not known, but there are several possible
explanations. One possibility is that a combination of
tryptases in the purified preparations is responsible for
PAR2 cleavage and activation. Another is that a posttranslational modification of PAR2 influences tryptase activation. PAR2 contains sites for N-linked glycosylation:
Asn30, 6 residues proximal to the cleavage and activation
site, and Asn222 in extracellular loop II. The potency with
which tryptase (but not trypsin) activates PAR2 is dramat-
ically increased by mutation Asn30, by enzymatic deglycosylation, or by expression of PAR2 in glycosylationdefective cells (62, 63). Thus glycosylation of the receptor
at a site close to the activation site markedly impairs the
capacity of tryptase to signal. Although the reason for this
finding is not known, glycosylation could impede access
of the amino terminus of PAR2 to the active site of
tryptase. Tryptase is a 134-kDa tetrameric protease in the
form of a flat ring that is composed of four monomers
whose active sites face the center of the ring (228). One
can speculate that a large, glycosylated structure may not
be accessible to the active sites. It will be important to
determine whether PAR2 is similarly glycosylated in tissues and to know if there are mechanisms that deglycosylate the receptor. However, can tryptase activate PAR2
in vivo? On balance, the evidence is in favor of an important role for tryptase in PAR2 activation under conditions
of inflammation and mast cell activation when large
amounts of tryptase are released close to PAR2 expressing cells. Thus injected tryptase has proinflammatory and
hyperalgesic actions in conscious mice that are not observed in PAR2-deficient animals (40, 296). Because
tryptase is a large and poorly diffusible protease, it is
likely then that tryptase signals in a paracrine manner to
cells that are in close proximity to mast cells, such as
sensory nerves that express PAR2, which participate in
inflammation and pain (267, 270, 296). Other proteases
from mast cells have not been shown to activate PARs,
although chymase can suppress thrombin signaling to
keratinocytes, suggesting that it disables PAR1 (254).
PROTEASE-ACTIVATED RECEPTORS
sitization of PAR2, suggesting that proteinase 3 is a PAR2
activator (292).
5. Cell-surface proteases
6. Nonmammalian proteases
An intriguing observation is that a number of nonmammalian proteases from mites, bacteria, and fungi
have been found to signal to mammalian cells by cleaving
and activating PARs. The dust mites Dermatophagoides
pteronyssinus and Dermatophagoides farinae produce a
series of proteases (cysteine proteases, trypsins, chymotrypsins, collagenases) that are allergens in the airway
epithelium. The cysteine and serine proteases Der P1 and
Der P9 stimulate cytokine release from human airway
epithelial cells and induce mobilization of Ca2⫹, effects
reminiscent of PAR activation (162). Indeed, Der P3 and
Der P9 can cleave fragments of PAR2 at the activation
site, and desensitization experiments suggest that the effects of these proteases on airway epithelial cells are
mediated in part by PAR2 (273). Bacterial proteases can
also signal through PARs. Porphyromonas gingivallis is a
major mediator of periodontitis in humans, and bacterial
arginine-specific gingipains-R (RgpB and HRgpA) have
been implicated in this disease. RgpB can activate PAR1and PAR2-transfected cells and signals to an oral epithelial cell line to induce release of the powerful proinflammatory cytokine interleukin-6 (178, 179). RgpB and
HRgpA can also signal to transfected cells expressing
PAR1 and PAR4, and low concentrations of both proteases
mobilize Ca2⫹ in platelets and induce aggregation with a
similar potency to thrombin (180). Cross-desensitization
studies and use of antibodies that block PAR cleavage and
activation suggest that the effects of gingipains-R on
platelets are mediated by PAR1 and PAR4. These interesting results reveal a novel mechanism by which bacteria
influence mammalian cells and could explain an emerging
link between periodontitis and cardiovascular disorders.
Certain fungi are also allergens in the airway, and proPhysiol Rev • VOL
teases from extracts of several species affect cytokine
production from airway epithelial cells (147). Some of
these effects are reminiscent of the activation of PAR2,
although the protease and receptor that are principally
responsible remain to be determined.
III. PROTEASE-ACTIVATED RECEPTORS:
TRANSDUCTION OF THE SIGNAL
A. Tethered Ligands Interact With Extracellular
Domains to Initiate Signal Transduction
Analyses of mutant and chimeric receptors and of
analogs of APs have allowed identification of the critical
residues of the tethered ligand domains that interact with
binding domains of the PARs and which are thus essential
for signal transduction. The consensus of these experiments is that conserved regions of extracellular loop II
and the amino terminus interact with the tethered ligands
of PAR1 and PAR2.
Interactions between the tethered ligand and PAR1
have been investigated by analysis of chimeras of human
and Xenopus PAR1 (105). The tethered ligands of human
(SFLLRN) and Xenopus (TRFIFD) PAR1 are specific for
their respective receptors and thus discriminate between
the receptors. Replacement of the extracellular amino
terminus and the second extracellular loop of the Xenopus receptor with the corresponding domains of human
PAR1 confers the chimera with selectivity for peptides
corresponding to the human tethered ligand. Moreover,
substitution of only two residues of Xenopus PAR1 with
corresponding residues of the human receptor (Phe87 for
Asn in the extracellular amino terminus and Glu260 for
Leu in the second extracellular loop) also confers selectivity for the human peptide. Thus these two residues in
the extracellular amino terminus and extracellular loop 2
are critical for interaction with the tethered ligand domain. Further mutational analysis of human PAR1 and
study of analogs of the PAR1 AP suggest that Glu260 of
extracellular loop 2 interacts with Arg5 of the tethered
ligand SFLLRN (201). Substitution of eight residues from
the second extracellular loop of the Xenopus receptor to
human PAR1 generates a receptor that is constitutively
active in the absence of ligand, suggesting that alteration
of the conformation of extracellular loop II is sufficient to
transduce a signal across the plasma membrane (202).
Interactions between the tethered ligand of PAR2 and
the cleaved receptor have been similarly examined by
studying chimeras of PAR1 and PAR2 (175). These studies
also reveal the importance of extracellular loop II for the
activation of PAR2. The important role of residues in
extracellular loop II of PAR2 for interaction with tethered
ligand peptides has also been revealed by the study of
mutant receptors and analogs of the PAR2 AP (4). An
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
Given that anchoring proteins serve as cofactors that
facilitate the capacity of certain proteases to activate
PARs, is it possible that proteases that are themselves
integral membrane proteins can activate PARs? One such
protease, membrane-type serine protease 1 (MT-SP1), has
been identified. MT-SP1 is a type II integral membrane
protein with an extracellular protease domain (276). Analysis of the substrate specificity of MT-SP1 suggests PAR2
as a potential substrate, and both MT-SP1 and PAR2 are
coexpressed at the surface of certain cell types (e.g., PC-3
cells). Indeed, solubilized MT-SP1 signals to oocytes expressing PAR2, but not to cells expressing PAR1, PAR3, or
PAR4. It remains to be determined if the membrane-bound
MT-SP1 can activate PAR2 under more physiological circumstances.
589
590
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
acidic region (PEE) that is just distal to a highly conserved domain (CHDVL) makes an important contribution to determining the selectivity of PAR2 agonists.
B. Structure-Activity Relations of Tethered
Ligand Domains
Physiol Rev • VOL
C. PARs Activate Multiple Signaling Cascades
In common with most GPCRs, PARs couple to multiple signaling pathways and can thereby regulate many
cellular functions. The mechanisms of PAR1 and PAR2
signaling have been extensively investigated and reviewed (89, 182).
1. Concentration-response relationships
The catalytic nature of PAR activation means that
even a very low concentration of an enzyme would eventually cleave every receptor molecule at the cell surface.
How then do proteases generate graded responses in
cells? For neurotransmitter receptors, graded concentrations of agonists induce graded responses through differing degrees of receptor occupancy. For PARs, graded
responses depend on the rate of receptor cleavage (137).
Thus thrombin induces cumulative phosphoinositide hydrolysis that correlates with cumulative receptor cleavage, suggesting that each cleavage event generates a
“quantum” of phosphatidylinositol signal. This signaling is
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
The observation that synthetic peptides corresponding to the tethered ligand domain of PAR1 are agonists for
the receptor without the need for proteolysis had three
important consequences. First, it enabled the use of synthetic peptides as probes for PAR function, thereby avoiding the sole use of proteases, which have many biological
effects not related to PARs. These APs are now widely
used to investigate the physiological functions of PARs in
vitro and in vivo. Second, it permitted convenient structure activity studies of the tethered ligand domain through
functional analyses of synthetic peptides (187, 253). Such
studies have provided important information about critical residues of the tethered ligand and facilitated the
development of more selective agonists (123). Third, analogs of the APs have been used as templates for the
development of antagonists of PAR1 (9). However, there
are certain caveats to the use of APs as PAR agonists.
Unfortunately, APs are weak agonists compared with proteases, often requiring concentrations in excess of 1 ␮M
to activate most PARs. The low potency of APs is mostly
attributed 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. Because APs
are used at such high concentrations, they may have
nonspecific effects that are unrelated to PAR activation.
The use of control peptides, such as inactive scrambled
peptides or reversed sequences, is thus essential, although these too can have activity (296). In addition, APs
are not always specific for a particular PAR; for example,
the PAR1 AP SFLLRN also activates PAR2 (21). Finally,
there are no effective APs for PAR3.
The characteristics of PAR APs and their utility as
templates for design of antagonists have been recently
reviewed in detail (182) and are only summarized here.
The first described PAR1 AP was a 14-residue peptide
(305), although it was soon realized that the hexapeptide
SFLLRN was fully functional (253). Although SFLLRN is a
PAR1 agonist, it also activates PAR2 (21). However, replacement of Ser1 with Thr1 (corresponding to the sequence in Xenopus PAR1) generates an agonist (TFLLRN)
that is selective for PAR1 and that does not activate PAR2.
Analysis of analogs of SFLLRN in which individual residues are substituted for Ala (alanine scanning), together
with site-directed mutagenesis of the tethered ligand, indicates that critical residues within this domain include
Phe2, Leu4, and Arg5. Substitution of other residues can be
tolerated, depending on the type of substitution. For instance, changes in Ser1 maintain function provided the
amino group is maintained, whereas removal of the amino
group abolishes activity. Detailed structure-function analyses of analogs of the PAR1 AP, together with analyses by
NMR and other techniques, have facilitated the design of
penta- and tetrapeptides with enhanced potencies. For
example, H-Ala-(pF)Phe-Arg-Cha-hARg-Tyr-NH2 induces
platelet aggregation with an EC50 of 10 nM, and an iodinated form of this peptide can be used in binding assays
(100). A peptidomimetic antagonist of PAR1 has been
developed that is selective for PAR1 over the other PARs
(9). Such antagonists have been successfully used in nonhuman primates to suppress thrombus formation and vascular occlusion (88) and could be of promise to treat
human disease. Analyses of analogs of the PAR2 AP have
similarly identified the residues that are essential for biological activity. In general, the rat/mouse peptide SLIGRL is slightly more potent than the human agonist
SLIGKV (21). Analysis by alanine scanning indicates that
Leu2 and Arg5 are essential for activity. Replacing Ser1 or
Arg5 with Ala also reduces activity, whereas substitution
of Gly4 or Leu6 has only a slight effect on PAR2 activation.
The PAR4 AP GYPGKF is neither as efficacious as thrombin in activating PAR4 nor as potent as APs for PAR1 or
PAR2 in activating the corresponding receptors, and is
thus of limited use for probing receptor function. However, the analog AYPGKF is ⬃10-fold more potent than
GYPGKF and is as efficacious as thrombin in activating
PAR4 (98). AYPGKF is also relatively specific for PAR4.
This specificity depends on Tyr2, since replacement with
Phe generates an agonist of PAR1 and PAR4.
PROTEASE-ACTIVATED RECEPTORS
rapidly terminated despite the continued presence of an
exposed tethered ligand, by mechanisms of desensitization that are discussed in section IVB. Thus cells sense the
rate of receptor cleavage, which is related to the concentration of protease. In other words, high concentrations of
thrombin rapidly cleave many PAR1 molecules, permitting
the accumulation of inositol 1,4,5-trisphosphate (InsP3)
and signal transduction. Proteases that inefficiently
cleave PARs could not rapidly generate a sufficient quantity of second messenger for signal transduction to occur.
2. PAR1 signaling
important role in coupling since thrombin signaling in
fibroblasts and platelets that express PAR1 is attenuated
by the microinjection of Gq␣ antibodies (13, 16). Moreover, platelets derived from Gq␣-deficient mice exhibit
markedly diminished thrombin-induced aggregation and
degranulation (222). These mice have increased bleeding
times, probably due to impaired thrombin signaling to
platelets. Gq11␣ activates phospholipase C-␤1, leading to
generation of InsP3, which mobilizes intracellular Ca2⫹,
and diacylglycerol (DAG), which activates protein kinase
C (PKC). Thus thrombin-stimulated InsP3 generation and
Ca2⫹ mobilization are blunted in mutant fibroblasts that
express low levels of phospholipase C-␤1 but normal levels of the other phospholipases (99). Thrombin-induced
activation of phospholipase A2 and phospholipase D is
diminished in this cell line, suggesting that phospholipase
C-␤1 activates PKC-␣, which may be required for phospholipase A2 and phospholipase D activation. Together,
Ca2⫹ and PKC activate numerous pathways, including
Ca2⫹-regulated protein kinases and mitogen-activated
protein (MAP) kinases.
PAR1 also couples to G12␣ and G13␣ in platelets and
FIG.
5. Summary of PAR1 signal
transduction. PAR1 couples to Gi␣, G12/
13␣, and Gq11␣. Gi␣ inhibits adenylyl cyclase (AC) to reduce cAMP. G12/13␣ couples to guanine nucleotide exchange factors (GEF), resulting in activation of Rho,
Rho-kinase (ROK), and serum response
elements (SRE). Gq11␣ activates phospholipase C␤ (PLC␤) to generate inositol
trisphosphate, which mobilizes Ca2⫹, and
diacylglycerol (DAG), which activates
protein kinase C (PKC). PAR1 can activate the mitogen-activated protein kinase
cascade by transactivation of the EGF
receptor, through activation of PKC,
phosphatidylinositol 3-kinase (PI3K),
Pyk2, and other mechanisms. G␤␥ subunits couple PAR1 to other pathways,
such as activation of G proteins receptor
kinases (GRKs), potassium channels
(Ki), and nonreceptor tyrosine kinases
(TK). [Modified from Coughlin (68).]
Physiol Rev • VOL
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
Signal transduction commences with coupling of
PARs to heterotrimeric G proteins at the plasma membrane. Considerable progress has been made in identifying the subtypes of G proteins that interact with PAR1,
and the role of these subunits in signaling and physiological regulation has been studied in knockout mice (reviewed in Ref. 219). PAR1 interacts with several ␣-subunits, in particular Gq11␣, G12/13␣ and Gi␣, which accounts
for the pleotropic action of its ligands (Fig. 5). A major
pathway of coupling is through Gq␣ (131). Gq␣ plays an
591
592
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
Physiol Rev • VOL
which in turn phosphorylates MAP kinase (serine-threonine kinase). MAP kinases in turn regulate multiple substrates in the cytoplasm and nucleus. The MAPK ERK1/2
module plays a critical role in cell proliferation and differentiation. Most information about the regulation of this
module derives from study of tyrosine kinase receptors,
such as the epidermal growth factor (EGF) receptor. 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 adaptor protein Shc, which recruits the
Grb2-SOS complex to exchange GDP for GTP on p21ras.
This initiates a cascade of phosphorylation events: p21ras
phosphorylates the serine-threonine Raf-1 kinase, a MAP
kinase kinase kinase, which phosphorylates MEK1/2
(MAP kinase kinase), which in turn phosphorylates
ERK1/2 (MAP kinase). There are several mechanisms by
which PAR1 can couple to this pathway. In astrocytes,
PAR1 activation of ERK1/2 and proliferation depend on a
pertussis toxin-sensitive pathway mediated by G␤␥, PI
3-kinase and ras, and a pertussis toxin-insensitive pathway involving PKC and raf (310). Another mechanism
may involve transactivation of the EGF receptor. Indeed,
activation of PAR1 induces phosphorylation of the EGF
receptor in enterocytes, and PAR1-induced Cl⫺ secretion
is suppressed by inhibition of EGF receptor kinase (30).
Possible mechanisms of transactivation include activation of the Ca2⫹-dependent kinase Pyk-2, and PAR1 agonists activate Pyk-2 in endothelial cells (161). Additionally, some GPCRs activate matrix metalloproteases,
which induce shedding of EGF receptor ligands from the
cell surface (231). PAR1 can also activate the MAP kinase
p38 module in fibroblasts by a mechanism involving EGF
receptor transactivation by Src family kinases (246).
The observation that mitogenic effects of PAR1 activators require tyrosine kinase activity, yet PAR1 does not
possess intrinsic tyrosine kinase activity, led to the investigation of the role of the janus family of tyrosine kinases
(JAKs) in PAR1 signaling and mitogenesis. JAKs tyrosine
phosphorylate STAT proteins (signal transducers and activators of transcription), which dimerize and translocate
to the nucleus to direct gene transcription. In vascular
smooth muscle cells, thrombin activates JAK2, resulting
in nuclear translocation of the transcription factors
STAT2 and STAT3 (183). Inhibition of JAK2 suppresses
thrombin-induced ERK2 activity and proliferation, suggesting that JAK2 is upstream of the Ras/Raf/MEK/ERK
pathway.
3. PAR2 signaling
Less is known about the signaling mechanisms of
PAR2 than for PAR1. Activators of PAR2 stimulate generation of InsP3 and mobilization of Ca2⫹ in PAR2-transfected cell lines, enterocytes, keratinocytes, myocytes,
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
astrocytoma cells (10, 220). Thus thrombin stimulates
incorporation of a GTP analog into G12␣ and G13␣ in
platelets, and thrombin-stimulated DNA synthesis is
blocked by microinjection of antibodies to G12␣ in astrocytoma cells. G12/13␣ plays a major role in control of cell
shape and migration. Upon stimulation with thrombin,
platelets change from a discoid to a spheroid shape and
extrude pseudopodia, which is believed to be a prerequisite for full activation. Although platelets from Gq␣-deficient animals show impaired thrombin-induced aggregation and degranulation (222), they undergo normal shape
changes in response to thrombin (163). The effects of
thrombin on the shape of these platelets is mediated by
G12/13␣ (163). G12/13␣ interacts with Rho guanine-nucleotide exchange factors (GEFs), permitting Rho-mediated
control of cell shape and migration. For example, in platelets G12/13␣ mediate activation of Rho-kinase and myosin
light-chain kinase, which participate in thrombin-induced
shape changes. Fibroblasts from G13␣-deficient mice
show diminished migratory responses to thrombin, and
these animals also exhibit impaired ability of endothelial
cells to develop into an organized vascular system, resulting in intrauterine death (221). In human umbilical vein
endothelial cells, PAR1 activators induce stress fiber formation, accumulation of cortical actin, and cell rounding
(303). Inhibition of Rho partially attenuates thrombininduced cell rounding, whereas dominant-negative Rac
blocks the response to thrombin. Thus Rho is involved in
the maintenance of endothelial barrier function and Rac
participates in cytoskeletal remodeling by thrombin.
PAR1 also couples to Gi␣ proteins, which inhibit
adenylyl cyclase and suppress formation of cAMP. Activation of PAR1 in fibroblasts inhibits cAMP generation in
a pertussis toxin-sensitive fashion, suggesting involvement of a Gi-like protein (132). Expression of a mutant of
Gi-2 in CHO cells suppresses thrombin-stimulated arachidonic acid release, suggesting that Gi-2 couples PAR1 to
cytoplasmic phospholipase A2 (315).
G␤␥ subunits of heterotrimeric G proteins couple
PAR1 to may other pathways, notably activation of phosphatidylinositol (PI) 3-kinase. PI 3-kinase thus links PAR1
to changes in cytoskeletal structure, cell motility, survival, and mitogenesis. For instance, in astrocytes, the
effects of PAR1 agonists on activation of extracellular
signal response kinases (ERKs) 1/2 and proliferation are
strongly inhibited by wortmannin, which blocks PI 3-kinase (310).
There has been considerable interest in understanding the mechanisms by which PAR1 couples to the MAP
kinase cascades, given the important mitogenic role of
thrombin. Several MAP kinase “signaling modules” have
been characterized in mammalian cells (reviewed in Ref.
314) with a common organization: a MAP kinase kinase
kinase (a serine-threonine kinase) phosphorylates and
activates MAP kinase kinase (threonine-tyrosine kinase),
PROTEASE-ACTIVATED RECEPTORS
neurons, astrocytes and tumor cells (24, 66, 67, 167, 213,
214, 251, 290). Therefore, it is likely that PAR2 couples to
Gq␣. Indeed, PAR2 signaling is unaffected by pertussis
toxin in PAR2-transfected cells and in enterocytes, suggesting that PAR2 does not signal through Gi proteins
(86). In enterocytes and transfected epithelial cells, activation of PAR2 also stimulates arachidonic acid release
and rapid generation of prostaglandins E2 and F1␣, suggesting activation of phospholipase A2 and cyclooxygenase-1 (167). PAR2 activators also strongly activate MAP
kinases ERK1/2 and weakly stimulate the MAP kinase
p38, although c-jun amino-terminal kinase is not activated
(15, 86, 319).
593
Recent observations indicate that ␤-arrestins play a
major role in PAR2-induced activation of ERKs. ␤-Arrestins are cytosolic proteins that interact with activated
GPCRs at the plasma membrane and in the cytosol (see
sect. IV). Activation of PAR2 stimulates the assembly of a
MAP kinase signaling module, a large-molecular-weight
complex containing PAR2, ␤-arrestins, raf-1, and activated
ERK1/2, which forms at the plasma membrane or in early
endosomes (86) (Fig. 6C). The module retains activated
ERK1/2 in the cytosol where they cannot induce proliferation. However, in cells expressing a mutant PAR2, which
is unable to interact with ␤-arrestins, the module is unable
to form. In these cells PAR2 agonists activate ERK1/2 by
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
FIG. 6. Desensitization, downregulation, and trafficking of PARs. Details vary between PARs, as discussed in the
text. A: receptor cleavage (1) activates PARs, which couple to heterotrimeric G proteins (2). Activation induces
membrane translocation of GRKs (GRK3 and 5 for PAR1), which phosphorylate the receptor (3) and promote membrane
translocation and interaction with ␤-arrestins, which mediate uncoupling and desensitization (4). B: receptor cleavage
and interaction with ␤-arrestins (PAR2, not PAR1) (1) induces clathrin- and dynamin-dependent endocytosis (2) and
rab5a-mediated trafficking to early endosomes (3). PARs are then sorted to lysosomes (4) by a mechanism involving
sorting nexin 1 (SNX1 for PAR1). Resensitization requires mobilization of intracellular pools of PARs (5). The pools are
replenished by new PAR synthesis or by constitutive endocytosis of unactivated PARs (6). C: in the case of PAR2,
activation induces recruitment of ␤-arrestins that serve as molecular scaffolds to recruit and organize components of the
mitogen-activated protein kinase (MAPK) pathway to coated pits (1) or early endosomes (2). ␤-Arrestins interact with
PAR2, raf-1, MEK1/2, and activated ERK1/2 to specify the subcellular location and function of activated ERK1/2.
Physiol Rev • VOL
84 • APRIL 2004 •
www.prv.org
594
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
alternate pathways that allow nuclear translocation and
stimulation of proliferation. Therefore, ␤-arrestins are
molecular scaffolds that promote the formation of an
ERK1/2 module at the plasma membrane or in endosomes. The scaffold retains activated ERKs in these locations, where they may regulate cytosolic targets. PAR2
agonists also activate the NF␬B pathway in keratinocytes
and myocytes, emphasizing the importance of PAR2 in
inflammation (28, 145).
IV. PROTEASE-ACTIVATED RECEPTORS:
TERMINATION OF THE SIGNAL
Proteases that remove or destroy the tethered ligand
or cleave the binding domain in extracellular loop II
would generate receptors that are unresponsive to activating proteases. Many proteases can disable PARs in this
manner (Table 1). However, the same criteria that should
be applied to consideration of the physiological relevance
of activating proteases (see sect. IIF) must also be applied
to any potential inactivating enzymes. Disabling proteases
may serve to dampen signaling by activating proteases
and could thereby be an additional mechanism for terminating protease signaling. A portion of proteolytically activated PAR1 can recycle to the cell surface in some cell
types, although it is usually targeted for degradation
(129). Proteolytically activated and recycled PAR1 can
continue to signal at the cell surface (283). In this case,
proteases that cleave or remove the exposed tethered
ligand would arrest signaling.
Proteases from inflammatory cells, including neutrophils and mast cells, can cleave and disable PARs. Neutrophil cathepsin G, elastase, and proteinase 3 cleave
PAR1 to remove the activation site and thereby abolish
thrombin signaling (197, 236). PAR1 AP still signals to
cells exposed to these proteases, suggesting that the binding domain is preserved. Chymase, an abundant mast cell
protease, also renders keratinocytes unresponsive to
thrombin, suggesting that it too inactivates PAR1 (254).
Elastase and cathepsin G cleave PAR2 in transfected cells
and airway epithelial cell lines, removing amino-terminal
epitopes and thereby generating receptors that are unresponsive to trypsin but normally responsive to PAR2 AP
(95). However, elastase and cathepsin G can signal to
fibroblasts to induce release of interleukin-8 and monocyte chemoattractant protein 1, possibly by activation of
PAR2 (291). Cathepsin G and elastase also abolish signaling by thrombin to PAR3-transfected cells, and thus disable PAR3 (74). Interestingly, this process of inactivation
occurs without the loss of binding of monoclonal antibodies that recognize sites flanking the tethered ligand domain, suggesting inactivation does not involve removal of
Physiol Rev • VOL
B. PAR Signaling Is Attenuated
by Receptor Desensitization
Proteases activate PARs by an irreversible mechanism: cleavage exposes the tethered ligand domain that is
always available to interact with the cleaved receptor.
Thus activation would result in prolonged signaling unless there were efficient mechanisms to attenuate the
response. The principal mechanism that terminates signaling by PARs is broadly similar to the classical pathway
of desensitization that has been described in detail for
many other GPCRs, in particular rhodopsin and the ␤2-
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
A. Cell-Surface Proteolysis Can Disable PARs
the tethered ligand. The mechanism of this inactivation is
unknown but could involve cleavage of tethered ligand
binding domains.
Some proteases can cleave PARs at several sites,
including activation and disabling sites, and the net result
depends on the efficiency of cleavage at different locations. For example, although cathepsin G can cleave PAR1
at the activation site Arg41-Ser42, the major cleavage site is
at Phe55-Trp56, which removes the tethered ligand and
disables the receptor (197). Although trypsin has been
reported to activate PAR1, it also efficiently cleaves PAR1
at distal sites that would remove the tethered ligand domain. Indeed, in endothelial cells, trypsin inactivates
PAR1, generating a receptor that is unresponsive to
thrombin (200). In addition to cleavage at the activation
site (Arg34-Ser35), tryptase also cleaves PAR2 at the Lys41Val42 site, which could inactivate the receptor (196). In
the case of tryptase, the activating cleavage is more important since tryptase activates PAR2.
The consequences of exposing cells to proteases depends on the repertoire of PARs expressed by a cell and
whether the proteases activate or disable particular receptors. For example, cathepsin G disables PAR1 but
activates PAR4. Thus in human platelets, which express
PAR1 and PAR4, cathepsin G can induce aggregation
through PAR4 even though it disables PAR1 and thereby
prevents signaling by low concentrations of thrombin
(249). In contrast, in fibroblasts and endothelial cells that
do not express PAR4, cathepsin G abolishes thrombin
signaling through PAR1. In addition, proteolysis can impair the ability of PAR3 to act as a cofactor for PAR4 (74).
In murine platelets, PAR3 binds thrombin but does not
signal. Instead, it concentrates thrombin in the vicinity of
PAR4 and thus serves as cofactor for PAR4 signaling in
response to low concentrations of thrombin. Cathepsin G
does not cause aggregation of murine platelets but prevents aggregation to low concentrations of thrombin, indicating that cathepsin G abolishes this cofactor role of
PAR3. It is very likely that proteases that destroy other
nonreceptor cofactors, such as TF, would also have profound effects on PAR function.
PROTEASE-ACTIVATED RECEPTORS
1. PAR1
PAR1 activators trigger rapid phosphorylation of the
receptor in transfected cell lines (136, 302, 304). Both
GRKs and second messenger kinases could mediate this
phosphorylation. Although activation of PKC stimulates
PAR1 phosphorylation, thrombin-stimulated phosphorylation persists after inhibition or downregulation of PKC,
suggesting other mechanisms of agonist-induced phosphorylation (136). One principal mediator of agonist-stimulated phosphorylation of PAR1 is GRK3, since overexpression of GRK3 enhances agonist-induced PAR1 phosphorylation and suppresses thrombin signaling, whereas
GRK2 is considerably less effective. The importance of
GRK3 in desensitization of PAR1 has been confirmed in
vivo by observations in transgenic mice overexpressing
GRK3 in the myocardium (133). Although signaling of the
␤2-adrenergic receptor is unaltered in these animals, since
this receptor is mostly regulated by GRK2, thrombininduced signaling is markedly attenuated by overexpression of GRK3. However, the observation that signaling of
GRK3-insensitive mutants of PAR1 is still rapidly attenuated suggests the existence of additional mechanisms,
perhaps involving other GRKs (136). GRK5 mediates
PAR1 desensitization in endothelial cells, where GRK5
expression suppresses thrombin signaling and a dominant
negative GRK5 mutant prolongs thrombin signals (278).
Physiol Rev • VOL
␤-Arrestins are cofactors for GRKs; they interact with
GRK-phosphorylated GPCRs at the cell surface and disrupt their association with heterotrimeric G proteins to
terminate the signal. ␤-Arrestins also mediate desensitization of PAR1 (225). Desensitization of PAR1 is markedly
diminished in embryonic fibroblasts prepared from mice
lacking both ␤-arrestin-1 and -2, and also in fibroblasts
lacking only ␤-arrestin-1, indicating a major role of ␤-arrestin-1 in desensitization of PAR1 in fibroblasts. This
observation is the first to show a differential role of
␤-arrestin isoforms in desensitization of a GPCR.
Several observations suggest that modifications of
the tethered ligand domain of cleaved PAR1 may also
contribute to desensitization. One mechanism may involve cleavage of the tethered ligand domain. Exposure of
platelets to PAR1 AP induces release of a PAR1 fragment
by a process inhibited by soybean trypsin inhibitor, suggesting that a trypsinlike enzyme cleaves activated PAR1
(223). In rat astrocytes, thrombin exposure strongly desensitizes Ca2⫹ mobilization to a second exposure to
thrombin and PAR1 AP (289). In contrast, exposure to
PAR1 AP desensitizes the response to subsequent challenge with thrombin but not to AP. These findings led to
the hypothesis that after activation with AP, the tethered
ligand domain is proteolytically destroyed, rendering the
receptor unresponsive to thrombin but not AP. Two observations support this hypothesis. First, as predicted,
exposure to thermolysin, which removes the tethered
ligand, generates a receptor that responds to PAR1 AP but
not to thrombin. Second, treatment with soybean trypsin
inhibitor enhances responses to thrombin and PAR1 AP,
suggesting the existence of an exogenous protease that
contributes to desensitization. The identity of the protease remains to be determined. The potential importance
of modifications of the exposed tethered ligand is also
illustrated by a study of PAR1 signaling in Sf9 insect cells.
PAR1 signals are prolonged in these cells, even after
washout of agonist (44). However, treatment with thermolysin, which removes the tethered ligand, attenuates
the otherwise sustained Ca2⫹ response.
2. PAR2
Responses to PAR2 activators are rapidly attenuated
and strongly desensitized to repeated exposure, indicating the existence of powerful mechanisms that control
signaling, which have not been fully characterized. PKC
appears to play an important role in regulating PAR2 (23,
86). Thus activation of PKC with phorbol esters abolishes
PAR2-induced Ca2⫹ signaling in transfected cell lines and
enterocytes that naturally express this receptor, whereas
PKC inhibitors magnify responses to PAR2 activators.
Mutation of a putative PKC site in the carboxy tail
(ST336/6 3 A) renders the receptor unresponsive to inhibition by phorbol esters (86). This receptor is also de-
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
adrenergic receptor (reviewed in Refs. 22, 181). The general mechanism of desensitization of these receptors is as
follows (Fig. 6A). Ligand occupation of the GPCR induces
the translocation of members of the family of G protein
receptor kinases (GRKs) from the cytosol to the activated
receptor at the cell surface. GRKs are serine-threonine
kinases that phosphorylate activated GPCRs, usually
within the carboxy terminus or third intracellular loop.
Phosphorylation triggers the membrane translocation of
arrestins, which interact with the phosphorylated GPCR,
disrupt association with heterotrimeric G proteins, and
thereby terminate signal transduction. Additional mechanisms may involve phosphorylation by second messenger
kinases, which also terminate signaling. However, there
remain many critical aspects of desensitization of GPCRs,
including PARs, that are unexplored. In many cases, the
GRK and arrestin isoforms that are coexpressed with
receptors in question and which mediate desensitization
to physiological stimuli are unknown. Most information
about desensitization derives from studies of transfected
receptors in cell lines; although genetically modified animals that are deficient in certain GRKs and arrestins are
available, comparatively little is known about mechanisms of desensitization in vivo. Mechanisms of desensitization also vary between different PARs, probably due
to structural differences, especially in the intracellular
loop III and carboxy terminus.
595
596
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
sensitization defective: Ca2⫹ responses to trypsin and
PAR2 AP are exaggerated, and there is diminished desensitization to repeated stimulation. The role of GRKs and
␤-arrestins in PAR2 desensitization is unknown.
3. PAR3 and PAR4
C. PARs Are Downregulated
by Intracellular Proteases
In addition to processes that regulate coupling of
receptors to signaling pathways, cells also determine their
responsiveness to agonists by regulating the levels of
receptors that are expressed at the plasma membrane and
which are thus accessible to agonists in the extracellular
fluid. The level of expression of receptors at the cell
surface is a balance between removal by endocytosis and
replenishment by recycling or mobilization of intracellular pools. Many receptors internalize after binding agonists. However, the fate of endocytosed receptors depends on postendocytic sorting, which varies from receptor to receptor. At one extreme, some GPCRs that are
activated in a reversible fashion by agonist binding, for
instance the neurokinin-1 receptor, efficiently recycle to
the plasma membrane to be reused by the cell (85). At the
other extreme, receptors such as the PARs, which, once
cleaved, cannot be reused by the cell, are destined for
intracellular degradation, which irrevocably terminates
signaling (Fig. 6B). For these receptors, recovery requires
synthesis or mobilization of new receptors. Of course, all
surface receptors have a finite life span and even recycling receptors are downregulated by intracellular degradation after long-term stimulation. The process of receptor downregulation is of considerable importance because defects can result in exaggerated signaling and
abnormal phenotypes, for example, transformation in the
case of EGF receptors that are not downregulated. PARs
are an ideal model to study these postendocytic sorting
events since they are invariably targeted for degradation
after activation. The molecular mechanisms and pathways of agonist-induced trafficking of PARs vary from
receptor to receptor and in different cells.
1. PAR1
PAR1 is rapidly and extensively internalized after
activation: stimuli trigger endocytosis of ⬎85% PAR1
Physiol Rev • VOL
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
Little is known about desensitization of PAR3 or
PAR4. PAR3 has a very short carboxy tail, with possible
implications for desensitization, since truncated receptors frequently exhibit diminished desensitization. PAR4
is not rapidly phosphorylated after activation, possibly
due to lack of sites for GRK phosphorylation in the carboxy tail, and in consequence, PAR4 slowly desensitizes
(263). The slow kinetics of desensitization may permit
PAR4 to signal in a sustained fashion in human platelets.
within 1 min in erythroleukemic and megakaryoblastic
cell lines, although endocytosis proceeds a little more
slowly in other cell types (129). The initial accumulation
of PAR1 into coated pits suggests that internalization
proceeds by a clathrin-mediated process. In support of
this suggestion, activated PAR1 colocalizes with clathrin
in PAR1-transfected fibroblasts, and disruption of clathrin
with hypertonic sucrose (causes abnormal clathrin polymerization) and monodansylcadaverine (disrupts imagination of clathrin-coated pits) markedly inhibits endocytosis of activated PAR1 (281). The GTPase dynamin mediates detachment of clathrin-coated pits, and expression
of a GTPase-defective dynamin mutant (K44E) also inhibits trafficking of activated PAR1. In addition to endocytosis of activated receptors, PAR1 also undergoes constitutive endocytosis in the absence of proteolytic activation.
However, whereas activated PAR1 is mostly destined for
lysosomal degradation (129), constitutively endocytosed
PAR1 recycles to the plasma membrane (261). Indeed,
there is a tonic endocytosis and recycling of PAR1 that
maintains receptor at the cell surface and in an intracellular pool. Distinct domains exist in the carboxy tail of
PAR1 that are required for stimulated and constitutive
endocytosis, suggesting that the two types are endocytosis proceeding by distinct mechanisms (262).
␤-Arrestins play a major role in endocytosis of many
GPCRs by serving as adaptor proteins that link GRKphosphorylated receptors to clathrin and AP2. However,
␤-arrestins play no role in stimulated endocytosis of
PAR1, which proceeds with normal kinetics in fibroblasts
lacking ␤-arrestin-1 and -2, whereas agonist-induced endocytosis of the ␤2-adrenergic receptor is inhibited in this
system (225). In addition, constitutive PAR1 trafficking
proceeds normally in fibroblasts that lack ␤-arrestins.
Different adaptor proteins, such as AP2, may participate
in endocytosis of PAR1.
Compared with our understanding of the molecular
mechanisms of endocytosis, less is known about the
mechanisms of postendocytic sorting of internalized
GPCRs, either to degradation or recycling pathways.
Once internalized, most activated PAR1 is targeted to
lysosomes for degradation. However, lysosomal sorting is
not complete because in some cells there is considerable
recycling (129). Progress has been made in defining the
molecular mechanisms that target PAR1 to lysosomes.
Sorting nexin-1 is a membrane-associated protein that
interacts with the carboxy tail of the EGF receptor and is
required for its downregulation. The yeast ortholog of
sorting nexin-1, Vps5p, participates in targeting hydrolases to the vacuole, the yeast equivalent of the lysosome.
In PAR1-transfected HeLa cells, agonist stimulation induces colocalization of PAR1 in endosomes with sorting
nexin-1 (311). Expression of a carboxy-terminal deletion
mutant of sorting nexin-1 has no effect on endocytosis of
activated PAR1, but induces its accumulation in early
PROTEASE-ACTIVATED RECEPTORS
2. PAR2
Like PAR1, activated PAR2 is endocytosed and trafficked to lysosomes (23, 90). Endocytosis of PAR2 proceeds by a clathrin-mediated mechanism since it is inhibited by blockade of clathrin function. Endocytosis of
PAR2 also requires dynamin, because GTPase defective
dynamin K44E inhibits endocytosis (242). In contrast to
PAR1, endocytosis of PAR2 requires ␤-arrestins. Thus activation of PAR2 induces the marked and rapid translocation of ␤-arrestin-1 from the cytosol to the plasma membrane, followed by a prolonged colocalization of PAR2
and ␤-arrestin-1 in early endosomes (86, 90). Eventually,
PAR2 and ␤-arrestin-1 are sorted into distinct compartments: PAR2 is targeted to lysosomes, and ␤-arrestin-1
returns to the cytosol. The importance of ␤-arrestins in
endocytosis of PAR2 is indicated by two observations.
First, expression of a dominant negative mutant of ␤-arrestin, ␤-arrestin318 – 419, which comprises the clathrin
binding domain, prevents endocytosis. Second, a PAR2
mutant, ST363/6 3 A, fails to internalize and shows impaired desensitization, because of its inability to interact
with ␤-arrestins. Thus PAR2 appears to belong to the class
B GPCRs that form high-affinity interactions with arrestins. ␤-Arrestins also play an important role in PAR2mediated activation of ERK1/2 as discussed in section
IIIC (86).
Progress has been made in defining the mechanisms
of intracellular trafficking of PAR2. The rab GTPases mediate many steps of vesicular trafficking of proteins to and
from the plasma membrane and between organelles.
Rab5a colocalizes with PAR2 in early endosomes within
minutes of activation (242). Expression of a GTPase defective mutant, dominant negative rab5aS34N, disrupts
early endosomes and inhibits endocytosis of activated
PAR2. Thus rab5a mediates distal steps in endocytic trafficking of PAR2 from clathrin-coated pits to early endoPhysiol Rev • VOL
somes. Once internalized, PAR2 translocates to lysosomes
and is degraded, as determined by Western blotting (Bunnett, unpublished observation). Although the molecular
mechanisms that target PAR2 for degradation are unknown, PAR2 is extensively ubiquitinated after activation,
determined by receptor immunoprecipitation and Western blotting for ubiquitin (Bunnett, unpublished observation), and ubiquitination of some other GPCRs is a prerequisite for degradation (185).
D. Sustained Signaling by Proteases Requires
Mobilization and Synthesis of New Receptors
PARs are activated by an irreversible mechanism,
and once cleaved, most activated PARs are destined for
lysosomal degradation. Thus PARs are “one-shot” receptors, and sustained signaling requires the mobilization of
intracellular stores of intact receptors or the synthesis of
new receptors. The importance of these mechanisms depends on whether cells possess prominent intracellular
stores of PAR1. In megakaryoblasts, where most receptors are rapidly internalized after thrombin cleavage, recovery is a slow process that depends on the synthesis of
new receptors (26, 129). In contrast, resensitization of
responses to PAR1 activators in endothelial cells and
fibroblasts depends on mobilization of the prominent
stores of PAR1 in the Golgi apparatus (121, 125). Platelets,
which lack both the ability to synthesize new receptors
and prominent intracellular pools, are unable to repopulate the plasma membrane with new receptors after
thrombin exposure and thus do not recover responsiveness to thrombin. Because platelets are only required to
respond once to thrombin, by aggregation, the lack of a
robust system of resensitization does not impair their
function. Moreover, human platelets also express PAR4,
which facilitates prolonged signaling by thrombin.
Similar mechanisms account for the resensitization
of PAR2. There are prominent intracellular pools of PAR2
in many cells, including enterocytes (23). Resensitization
of responses to trypsin and the replenishment of the cell
surface with intact PAR2 is minimally affected by cycloheximide but attenuated by disruption of the pools with
brefeldin A, suggesting that resensitization requires mobilization of pools. Similar mechanisms of PAR2 resensitization have been observed in vascular tissues in vitro,
suggesting that these are widespread mechanisms (113).
The role of rab GTPases in trafficking of PAR2 from
Golgi stores to the cell surface has been examined.
Rab11a colocalizes in the Golgi apparatus with PAR2, and
PAR2 activators stimulate redistribution of rab11a into
vesicles containing PAR2 that migrate to the cell surface
(242). Expression of GTPase defective rab11aS25N
causes retention of PAR2 in the Golgi apparatus, and
overexpression of wild-type rab11a accelerates both re-
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
endosomes and inhibits its degradation of lysosomes.
Thus sorting nexin-1 participates in sorting of PAR1 from
early endosomes to lysosomes.
The domains of PAR1 that specify targeting to lysosomes are not fully defined, although substitution of the
carboxy tail of PAR1 with the equivalent domain of the
neurokinin-1 receptor (which recycles) alters PAR1 trafficking to the recycling pathway (282, 283). This recycled
receptor appears to be tonically active, which emphasizes
the importance of lysosomal degradation as a mechanism
that unequivocally terminates PAR1 signaling. Conversely,
the neurokinin-1 receptor with the carboxy tail of PAR1 is
targeted to lysosomes. Thus carboxy-terminal domains of
PAR1 are important for lysosomal trafficking. This lysosomal trafficking of PAR1 (and PAR2) proceeds whether
the receptor is activated by proteolysis or by AP, and thus
is not unique for the cleaved receptor.
597
598
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
covery of PAR2 at the cell surface and resensitization of
PAR2 signaling. Thus rab11a plays an important role in
PAR2 resensitization. Of interest, rab5a promotes and
rab5aS34N impedes resensitization of cells to trypsin.
Thus rab5a is required for PAR2 endocytosis and resensitization, whereas rab11a contributes to trafficking of
PAR2 from the Golgi apparatus to the plasma membrane.
V. PROTEASE-ACTIVATED RECEPTORS:
CONTRIBUTION TO PHYSIOLOGICAL
AND PATHOPHYSIOLOGICAL
CONTROL MECHANISMS
FIG. 7. A summary of the potential physiological and pathophysiological roles of proteases and PARs in different
cell types. Signaling depends on the availability of proteases, which can activate or disable PARs, protease inhibitors,
PARs, and protease cofactors and anchoring proteins. Proteases signal to a wide variety of cells to regulate critically
important biological processes, with implications for physiological regulation and disease.
Physiol Rev • VOL
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
Although much is known about the potential functions of PARs (Fig. 7), there remain substantial obstacles
to understanding the role of proteases and PARs in physiology and disease. First, the system is complex: protease
signaling depends on availability of activating or disabling
proteases, protease inhibitors, and cofactors or anchoring
proteins. Understanding the system requires knowledge
of all of these components, and in many situations this
information is lacking. Second, most studies rely on administration of activators (proteases or APs). Although
such studies provide information about the potential role
of PARs, they do not provide direct information about
their function in physiology or diseases states. Proteases
and certain APs are not selective activators of PARs, and
APs are agonists at only very high concentrations (usually
⬎1 ␮M). Peptides at high concentrations can have biological effects that are unrelated to receptor occupancy. For
example, inactive analogs of PAR2 AP cause edema upon
intraplantar injection (296). Third, selective antagonists
of PARs are not widely available. Although antagonists
have been developed for PAR1 (9), the intramolecular
mechanism of activation of PARs may hamper the identification of antagonists by traditional methods involving
screening of large chemical libraries. There are recent
reports of peptide-based antagonists of PAR1 and PAR4
PROTEASE-ACTIVATED RECEPTORS
A. Protease Signaling in the Circulatory
and Cardiovascular Systems
The contributions of proteases and their receptors to
regulation the cardiovascular system have been recently
reviewed (50, 68, 226, 240).
1. Proteases
As discussed in section III, coagulation and anticoagulation factors, in particular thrombin, FVIIa, FXa, and
APC, can regulate cells by cleaving PARs. Proteases released from inflammatory cells within the circulation and
in tissues can also activate PARs in the circulatory and
cardiovascular systems. For example, cathepsin G from
neutrophils is a potential activator of PAR4 on platelets
(249). Endothelial cells also express trypsins (169), which
could activate PAR2 and PAR4, although almost nothing is
known about the function of endothelial trypsins.
2. PARs
PARs are expressed by circulating cells as well as by
endothelial and vascular smooth muscle cells. Platelets
Physiol Rev • VOL
express several receptors for thrombin, although the combination varies between species: human platelets express
PAR1 and PAR4 (142), murine platelets express PAR3 and
PAR4 (143), and guinea pig platelets express PAR1, PAR3,
and PAR4 (8). Neutrophils also express PAR2 (128). Human umbilical vein endothelial cells express PAR1, PAR2,
and PAR3 (194, 218), and vascular smooth muscle cells
express PAR1 and PAR2 (79).
3. PAR signaling to circulating cells
Platelets are circulating cells that are critical mediators of coagulation. Thrombin cleaves PARs on platelets
to cause aggregation, granular secretion, and induction of
procoagulant activity, which are essential for hemostasis.
Thrombin signaling to platelets is enhanced by interaction
with surface proteins such as glycoprotein I␤␣ (84) or by
direct binding to PAR1 or PAR3, which possess hirudinlike thrombin-binding domains (135, 306). Thrombin signals by activating PAR1 and PAR4 in human platelets (142)
and PAR4 alone in murine platelets, since PAR3 does not
signal in murine platelets (199). What is the relative importance of PAR1 and PAR4 in human platelets? In isolated human platelets, antagonism of PAR1 blocks responses to low concentrations of thrombin, and antagonism of both receptors inhibits responses to low and high
thrombin concentrations (142). Activation of PAR1 with
AP is as efficacious as thrombin in inducing aggregation,
Ca2⫹ influx, and development of procoagulant activity in
human platelets, whereas PAR4 AP is less efficacious (7).
Although these results could indicate that PAR1 is the
primary thrombin receptor on human platelets, with PAR4
functioning in reserve, they may also reflect a generally
lower efficacy of PAR4 AP as an agonist of that receptor.
PAR4 may also contribute to sustained actions of thrombin on platelets and thereby mediate the late phase of
platelet activation (70). In addition, PAR4 could act as a
platelet receptor for other proteases, such as cathepsin G
(249). In murine platelets, which express PAR3 and PAR4,
PAR3 does not signal but instead anchors thrombin to the
cell surface by its hirudin-like domain and thereby facilitates thrombin cleavage of PAR4 (199, see sect. IID2).
The contributions of PARs to coagulation have also
been investigated in vivo by use of antagonists to PAR1. In
cynomologus monkeys, whose platelets express PAR1
and PAR4, an antagonist of PAR1 inhibits thrombus formation and vessel occlusion in a model of electrolytic
injury of the carotid artery (88). These studies indicate
that PAR1 plays a predominant role in platelet regulation
in this species. However, the importance of a dual receptor system on murine platelets in vivo is illustrated by the
observation that deletion of PAR3 or PAR4 results in
prolonged bleeding time and protection from ferric chloride-induced thrombosis of mesenteric arterioles and
thromboplastin-induced pulmonary embolism (313). The
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
(termed pepducins) that mimic intracellular domains of
these receptors, but they have yet to be widely used (71,
72). Although it is generally more straightforward to develop drugs that inhibit proteases, it can be difficult to
attain absolute selectivity between related enzymes.
In the absence of selective activators and antagonists, the best current approach to specifically determine
the role of PARs in vivo is to study genetically modified
animals. Mice have been developed that lack all known
PARs (64, 83, 143, 177, 255, 313), and by selective crossing
it will be possible to breed animals that lack several
receptors. Transgenic animals have also been identified
that overexpress PAR2 (255). However, the use of genetically modified mice is also fraught with problems. Effects
of deletion on embryonic survival and fertility impede the
breeding of animals for experimentation: approximately
one-half of PAR1 knockout mice die before birth (64, 83).
Compensatory alterations in the expression or activity of
other PARs can occur in knockout animals (77), but in
general these possibilities have not been thoroughly
tested. There are also major differences in biological responses between different strains of mice, and it is essential to study wild-type and knockout animals in the identical genetic background, requiring the use of littermates
or extensive back-crossing. Finally, many experimental
models using mice are very imperfect models of human
disease, requiring caution in extrapolating observations in
mice to humans.
With these caveats in mind, we will briefly discuss
the roles of proteases and their receptors in several systems.
599
600
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
role of PAR4 in vivo has also been examined by use of
novel antagonists, the pepducins (71). Peptides derived
from the third intracellular loop of PAR1 and PAR4 with
attached palmitate groups are potent inhibitors of thrombin-mediated aggregation of human platelets. Treatment
of mice with a PAR4 pepducin prolongs bleeding time and
protects against systemic platelet activation, consistent
with the phenotype of PAR4-deficient mice (72). Thus
PAR4 plays an important role in thrombin signaling in
mice.
4. PAR signaling to the vessels
Physiol Rev • VOL
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
Proteases can signal to endothelial cells and vascular
smooth muscle cells to control resistance and blood flow,
extravasation of plasma proteins and granulocytes, and
angiogenesis.
The effects of PAR activators on the contractile state
of blood vessels has been extensively examined by classical organ bath pharmacology. The net effect of PAR
activators depends on the receptor, the vessel, and
whether the effect is mediated by release of agents from
the endothelium of by direct actions on vascular smooth
muscle. Thrombin and selective PAR1 agonists invariably
cause relaxation of precontracted blood vessels, including large conduit-like vessels, for example, human pulmonary artery (116), porcine coronary artery (114), and rat
aorta (184), and also of smaller resistance vessels, including human and porcine intramyocardial arteries (117).
The consensus of these studies is that PAR1-induced relaxation is prevented by removal of the endothelium and
is thus mediated by the secretion of a factor from endothelial cells that relaxes vascular smooth muscle. Relaxation is frequently suppressed by inhibition of nitric oxide
(NO) synthase and is thus mediated in part by the release
of NO from endothelial cells. However, there is also an
NO-independent mechanism. In porcine coronary artery,
this NO-independent mechanism involves the release of
an endothelial-derived hyperpolarizing factor (114). In human pulmonary arteries, products of cyclooxygenase also
contribute to relaxation (116). PAR1 activators can also
contract certain vessels by endothelium-dependent and
-independent mechanisms (277).
Activators of PAR2 also relax vascular tissues. Thus
trypsin and PAR2 AP cause relaxation of a wide range of
vessels from several species, including rat pulmonary artery (243) and human and porcine coronary and intramyocardial arteries (117). PAR2-mediated relaxation depends
on the presence of the endothelium and involves both
NO-dependent and -independent mechanisms. PAR2 AP
also induces an endothelium-dependent contraction of,
for example, rat aorta and renal, mesenteric, and pulmonary arteries (243). However, this contraction may involve a subtype of PAR2 in view of the different potencies
with which analogs of the PAR2 AP induce contraction
versus relaxation. The existence of a subtype receptor is
also suggested by use of a “sandwich assay,” in which a
segment of human umbilical vein is sandwiched next to a
segment of endothelium-denuded rat aorta that serves as
a “reporter” for the release of vasoactive agents from the
human tissue (248). PAR2 AP induces release of an unknown factor from human umbilical vein that stimulates
contraction of a rat aorta. Notably, trypsin does not have
this effect, and the mechanism is thus independent of
PAR2.
The effects of PAR activators on cardiovascular responses have been investigated in intact animals. In mice,
intravenous injection of PAR1-selective AP causes a rapid
and sustained hypotension, which is not NO mediated
(47). Similarly, in rats and mice, PAR2 activators cause
hypotension by NO-dependent and -independent mechanisms, which can be followed by a reflex hypertension
(52). The effects of PAR2 activators on blood flow have
recently been investigated in humans. Administration of
PAR2 AP causes dilation of resistance vessels in the arm
with an elevated blood flow, by NO- and prostaglandindependent mechanisms (241).
What are the physiological or pathophysiological implications of regulation of blood flow by PARs? Under
normal circumstances it is unlikely that PARs regulate
blood flow; indeed, there are no obvious abnormalities in
the circulatory system in mice lacking PARs. However,
PARs may play an important role during inflammation and
sepsis, when there are elevated levels of proteases from
inflammatory cells, the coagulation cascade and other
sources, together with an elevated expression of PARs
themselves. Exposure of endothelial cells in culture to
inflammatory mediators such as TNF-␣ and lipopolysaccharide induces a selective upregulation of PAR2 but not
PAR1 (216). Administration of endotoxin to rats also results in an upregulation of PAR2 in the vascular system
that is accompanied by an increased hypotensive response to PAR2 activators (53). Similarly, exposure of
segments of human coronary artery to interleukin-1␣ and
TNF-␣ upregulates PAR2 and PAR4 and exacerbates
PAR2- and PAR4-mediated relaxation of tissues (115).
Inflammation is accompanied by marked alterations
in the permeability of blood vessels to plasma proteins
and cells. Many inflammatory signals trigger the formation of gaps between endothelial cells of postcapillary
venules, resulting in plasma extravasation and edema;
these stimuli can also induce expression of adhesion molecules that promote the firm adhesion and ultimate infiltration of granulocytes into tissues. Proteases from the
coagulation cascade and from inflammatory cells can trigger these events. The effects of thrombin and PAR1 on
extravasation of plasma proteins have been studied in
cultured endothelial cells and in the intact animal. Thrombin stimulates contraction of endothelial cells from the
human umbilical vein, which results in gap formation and
PROTEASE-ACTIVATED RECEPTORS
Physiol Rev • VOL
vascular matrix after injury (48). Activation of PAR2 also
stimulates proliferation of endothelial cells (194). In a
murine model of limb ischemia, trypsin and PAR2 AP
promote angiogenesis and accelerate homodynamic recovery of the limb, suggesting a role for PAR2 in angiogenesis (191).
B. Protease Signaling to the Nervous System
The role of proteases and PARs in the nervous system is a topic of recent reviews (295, 308).
1. Proteases
Protease activators of neuronal PARs may derive
from the circulation, from inflammatory cells that are
resident in or recruited to neural tissues, or from neurons
or astrocytes. Coagulation cascade proteases could signal
to the nervous system during trauma and inflammation
when there is elevated vascular permeability. However,
these proteases are also expressed within the nervous
system, although at low levels. Prothrombin transcripts
are present in certain regions of the brain that also express PAR1 (91, 312). For example, in the olfactory bulb
prothrombin and PAR1 mRNA are coexpressed in the
mitral and granular cell layers. Elsewhere in the olfactory
bulb and neocortex, the diffuse expression of prothrombin contrasts with the more defined expression of PAR1.
Nevertheless, the potential exists for neuronally derived
thrombin to activate PAR1 in the nervous system. Proteases from inflammatory cells may also signal to the
nervous system. Granzyme A is released from CD4⫹ cytotoxic lymphocytes that infiltrate the brain during experimental allergic encephalomyelitis (272). Granzyme A
cleaves a fragment of PAR1 at the activation site, and
granzyme A signaling to astrocytes and neurons is inhibited by a PAR1 blocking antibody, suggesting that granzyme A activates PAR1. Neural tissues also express a large
number of proteases that potentially activate PARs (252).
The biological effects of proteases are tempered by
protease inhibitors, and neuronal tissues are replete with
an array of inhibitors for proteases that activate PARs.
The serpin protease nexin-1 is the most efficient endogenous inhibitor of thrombin. Nexin-1 mRNA is expressed in
the brain where it is found in glial cells, synapses, and at
high levels in capillaries and vascular smooth muscle in
the human cortex (102). Thrombomodulin is a transmembrane glycoprotein that can form complexes with proteases and thereby regulates their activity at the cell
surface. Thrombomodulin is expressed by astrocytes in
culture, and expression of thrombomodulin is upregulated during trauma and consequent astrogliosis, as well
as by activation of PAR1 (230). The upregulation of thrombomodulin by thrombin may thus serve to control activation of PAR1 on astrocytes.
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
increased permeability of confluent cells (174). The intraplantar injection of thrombin and selective PAR1 AP
causes edema formation in the rat paw, indicative of
plasma extravasation (299). This effect is partly dependent on degranulation of mast cells and release of serotonin and histamine, and partly due to activation of PAR1
on the peripheral endings of primary spinal afferent neurons, and release of SP (87). By use of intravital microscopy it has been shown that thrombin stimulates leukocyte rolling and adhesion when applied topically to postcapillary venules of the rat mesentery (297). However,
this response is neither prevented by a PAR1 antagonist
nor reproduced by PAR1-selective AP, but may be mediated by PAR4 since PAR4 AP also induces the response. In
a similar manner, activation of PAR2 induces edema of the
paw and also causes infiltration of granulocytes (298). In
this case the edema is mediated in large part by the
release of neuropeptides from sensory nerves and is thus
dependent on a neurogenic mechanism (270). However,
PAR2-induced infiltration of granulocytes is not mediated
by the nervous system, but probably by direct effects of
PAR2 activators on endothelial cells. Activation of PAR2
also stimulates leukocyte rolling, adhesion, and extravasation in postcapillary venules of the rat mesentery, by a
mechanism that depends on release of platelet activating
factor (294). Similarly, administration of PAR2 AP to the
cremaster muscle of mice induces leukocyte rolling and
adherence to venules (177). There is a modest decrease in
the surgically induced rolling of neutrophils that is observed in PAR2-deficient mice, suggesting involvement of
PAR2 in inflammation.
Activation of PAR1 by thrombin and AP stimulates
proliferation of endothelial (194) and vascular smooth
muscle cells (189). Activation of PAR4 similarly induces
proliferation of vascular smooth muscle cells, suggesting
that PAR1 and PAR4 may mediate the effects of thrombin
(27). The proangiogenic effects of thrombin involve interactions with vascular endothelial growth factor (VEGF),
itself a potent angiogenic factor. PAR1 activators upregulate expression of VEGF receptors (KDR and flt-1) by
endothelial cells and potentiate the mitogenic effects of
VEGF on endothelial cells (285). VEGF itself also promotes thrombin generation. These results suggest the
existence of a mechanism of amplification of the angiogenic response involving PAR1 and VEGF receptors. In
view of the proangiogenic effects of thrombin, it is not
surprising that thrombin and PAR1 may play a role in
vascular development. Deletion of prothrombin results in
abnormalities of the vasculature of the yolk sac, and
deletion of PAR1 results in partial embryonic lethality that
is associated with bleeding abnormalities and which can
be rescued by overexpression of PAR1 in endothelial cells
(110). The role of PAR1 in responses to mechanical damage has also been evaluated in the mouse, which indicates
involvement of PAR1 in regulation of the formation of the
601
602
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
2. PARs
3. PAR signaling in the nervous system
The regulation of [Ca2⫹]i in neuronal cells is of fundamental importance in the control of excitability, release
of neurotransmitters, and survival, and the signaling of
proteases in neurons and astrocytes is commonly studied
by measuring generation of second messengers, such as
Ca2⫹. All four PARs are expressed in astrocytes (309).
Thrombin and PAR1 AP increase [Ca2⫹]i in a glioma cell
line and in astrocytes from the rat brain (287, 288). The
nature of the response depends on the duration of the
stimulus: brief exposure triggers a rapid, transient increase in [Ca2⫹]i, whereas continued exposure induces
oscillations or a sustained increase that are attributable to
an influx of extracellular Ca2⫹. The influx of extracellular
Ca2⫹ and the refilling of internal stores of Ca2⫹, which are
both essential for oscillations in [Ca2⫹]i, depend on actiPhysiol Rev • VOL
4. Morphology
Innervation depends on the control of neurite outgrowth, and there is great interest in understanding the
factors that control this process. Proteases and protease
inhibitors regulate the outgrowth of neurites. Neuroblastoma cells and certain primary cultures of neurons rapidly
extend neurites when switched from serum-containing to
serum-free culture medium. Thrombin inhibits neurite
sprouting in serum-free medium (112). Conversely, neurite outgrowth in the presence of serum is stimulated by
the thrombin inhibitors protease nexin-1 and hirudin,
probably due to inhibition of thrombin in serum. Thrombin similarly inhibits the outgrowth of neurites from dorsal root ganglia (106). Astrocytes contribute to the bloodbrain barrier, and there is considerable interest in understanding the control of astrocytes morphology. Protease
and protease inhibitors regulate the morphology of astrocytes (14, 39). When cultured in serum, astrocytes adopt a
flattened, fibroblast-like morphology, whereas in serumfree conditions they become stellate and project long
processes. Thrombin mimics the effects of serum,
whereas protease nexin-1 reverses the effect of thrombin.
These effects of thrombin on the morphology of neurons
and astrocytes are mimicked by PAR1 AP and are thus
likely attributable to activation of PAR1. Rho GTPases
regulate the actin-based cytoskeleton and thereby participate in the control of cell shape and migration. Indeed,
thrombin-stimulated neurite retraction and rounding of
neurons is blocked by treatment of cells with Clostridium
botulinum toxin C3 exoenzyme, which ADP-ribosylates
and thereby inactivates small Rho GTP-binding proteins,
implicating Rho in this action of thrombin and PAR1
(138).
5. Survival and proliferation
Proteases and PARs have important effects on the
survival of neurons and astrocytes, with major implications for neuronal functions during trauma and inflammation. The effects of thrombin on survival of neurons and
astrocytes depend on the concentration of activators and
the duration of exposure. Low concentrations of PAR1
activators are protective. Thrombin (⬍10 nM, 48 h) and
PAR1 AP markedly protect cultured astrocytes and hypo-
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
Expression of PAR1 mRNA in the rat brain is higher
at birth compared with postnatal day 28, and levels increase again in adulthood (206, 207). PAR1 is widely expressed throughout the rat brain, with higher levels of
expression in particular regions, where it is likely to be
present in neurochemically and functionally distinct cell
types (312). Thus PAR1 mRNA is expressed in the neocortex, cingulate/retrosplenial cortex, subiculum, nuclei
within the hypothalamus and thalamus, and in discrete
layers of the hippocampus, cerebellum, and olfactory
bulb. In the hypothalamus, PAR1 is present in dopaminergic neurons, but it is likely to be present in GABAergic
granule cells in the olfactory bulb and in glutaminergic
neurons of the anterior thalamus. PAR1 is also expressed
at low levels in glial cells throughout the brain, and by
ependymal cells of the choroid plexus and ventricular
lining where it could participate in control of transport of
cerebrospinal fluid. Within the spinal cord, PAR1 mRNA is
expressed in motoneurons and in preganglionic neurons
of the autonomic nervous system. PAR1 is also present in
the peripheral nervous system. In dorsal root ganglia,
PAR1 colocalizes in neurons with the neuropeptides SP
and calcitonin gene-related peptide (CGRP) (87). These
are primary spinal afferent neurons that are known to
participate in neurogenic inflammation and nociception.
PAR1 is present in more than half of the neurons in the
myenteric plexus of the guinea pig small intestine, suggesting a role for PAR1 in the control of intestinal
motility (67).
PAR2 is also expressed within the brain. Thus PAR2 is
found within neurons and glial cells of the hypothalamus
(79, 265). In the periphery, PAR2 is expressed in primary
spinal afferent neurons with SP and CGRP (270) and by a
large proportion of neurons in the myenteric and submucosal plexuses of the guinea pig and mouse small intestine
(67, 73, 235).
vation of protein kinase C by PAR1. The response of
astrocytes to thrombin may be mediated by two coexpressed receptors, PAR1 and PAR4 (148). Thus, in primary
cultures of human astrocytomas, thrombin as well as
PAR1 and PAR4 AP induce elevations in [Ca2⫹]i. Normal
brain astrocytes express PAR2 and respond to trypsin and
PAR2 AP (290). Selective agonists of PAR1 and PAR2 also
signal enteric neurons and primary spinal afferent neurons to increase [Ca2⫹]i, providing functional evidence for
expression of these receptors (67, 87, 270).
PROTEASE-ACTIVATED RECEPTORS
Physiol Rev • VOL
in these cells. Thus antagonists of PAR4 could be effective
therapies for neurotrauma.
7. Neurogenic inflammation and pain
A subpopulation of primary sensory neurons with
cell bodies in dorsal root, trigeminal, and vagal (nodose
and jugular) ganglia express the neuropeptides SP and
CGRP. These neurons have small, dark cell bodies with
unmyelinated (C) or thinly myelinated (A-␦) fibers and
participate in neurogenic inflammation and nociception.
The peripheral projections of these fibers detect thermal,
chemical, and high-threshold mechanical stimuli in many
tissues, including the skin, gastrointestinal tract, and airway. Stimulation results in the release of SP and CGRP
from endings of primary sensory neurons in peripheral
tissues, which initiates a variety of responses, collectively
referred to as “neurogenic inflammation,” and characterized by arteriolar vasodilatation, extravasation of plasma
proteins from postcapillary venules, and adhesion of leukocytes to the venular endothelium. These same stimuli
also generate action potentials that are transmitted centrally where they induce the release of neuropeptides
within the central nervous system, resulting in the central
transmission of nociceptive signals. Proteases that activate PAR1 and PAR2 can signal to these neurons in peripheral and perhaps central tissues to control inflammation and pain (11, 87, 270, 296) (Fig. 8).
The role of PAR2 in sensory neurons has been extensively investigated. PAR2 is expressed by ⬃65% of neurons in the dorsal root ganglia of the rat; 25–35% of these
PAR2-expressing neurons also express SP and CGRP, respectively, and are thus involved in inflammation and pain
(270). Trypsin, tryptase, and PAR2 AP signal to these
neurons to elevate [Ca2⫹]i and to stimulate secretion of
SP and CGRP from segments of the dorsal horn, bladder,
and atria, indicating that activation of PAR2 triggers peptide release from both the central and peripheral projections of these neurons. The intraplantar injection of PAR2
activators causes a sustained (⬎6 h) edema that is mediated by release of SP and CGRP from the peripheral
projections of sensory nerves. Activation of PAR2 also
induces hyperalgesia (149, 150, 296). Thus intraplantar
injection of subinflammatory doses of PAR2 AP causes
sustained (⬎24 h) mechanical and thermal hyperalgesia,
with increased expression of fos protein in spinal neurons, indicative of neuronal activation. Hyperalgesia is
not observed in mice lacking the neurokinin-1 (SP) receptor or preprotachykinin A, which encodes SP and neurokinin A. Neurokinin-1 receptor antagonists also prevent
PAR2-induced hyperalgesia, but they are effective only if
capable of penetrating the spinal cord or if delivered
intrathecally. Thus PAR2-mediated hyperalgesia depends
on the release of SP from the central projections of primary spinal afferent neurons. Together, these findings
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
thalamic neurons from death in response to environmental insults, including hypoglycemia and oxidative stress
(293). Similarly, thrombin (50 pM, 1 h) and PAR1 AP
protect hippocampal neurons in brain slices from death in
response to experimental ischemia (271). This protective
effect of thrombin is also observed in vivo. Thus brief
occlusion of the common carotid arteries of gerbils induces tolerance of hypothalamic CA1 neurons to a more
sustained ischemia. Intracerebral injection of the thrombin inhibitor hirudin prevents this protection, indicating
involvement of thrombin (271). Thrombin-induced protection of astrocytes from hypoglycemia is attenuated by
general inhibitors of serine/threonine kinases and by inhibition of Rho, suggesting involvement of the actin cytoskeleton (92). Another protease that activates PAR1,
APC, also protects the brain from ischemic injury by a
PAR1-dependent mechanism. Thus the protective actions
of APC in a model of ischemic brain injury in mice are
markedly attenuated by administration of a PAR1 blocking antibody (46).
However, activation of PAR1 also induce neurodegeneration. Prolonged exposure of neurons and astrocytes to high concentrations of thrombin (500 –750 nM)
causes apoptosis (93), and thrombin (100 nM, 1 h) decreases survival of hippocampal neurons in brain slices
after hypoglycemic and hypoxic stress (271). The same
kinase and Rho-dependent pathways that mediate protection also induce apoptosis, the difference resting on the
duration and extent of receptor activation. Moreover,
whereas low concentrations of thrombin induce transient
elevations in [Ca2⫹]i in hippocampal neurons, high concentrations induce large and sustained elevations, which
may well trigger apoptosis (271). Supporting this concept,
caspase-3 inhibitors prevent PAR1-induced death of motoneurons in the avian embryo (286). Some of the detrimental effects of thrombin and PAR1 APs could be mediated by the N-methyl-D-aspartate (NMDA) receptor system, which contributes to degeneration in the nervous
system. Thus PAR1 activation potentiates NMDA currents
in hippocampal neurons, raising the possibility that
thrombin could exacerbate glutamate-induced neurotoxicity (107). In addition to effects on survival, activation of
PAR1 also induces proliferation of astrocytes (229). The
dual protective and degenerative effects of PAR1 could
have implications for survival of neurons after injury:
immediately after trauma and generation of thrombin,
PAR1 may serve to protect the brain from damage; thereafter, PAR1 may have deleterious effects.
PAR4 may also contribute to inflammation in the
brain, with detrimental effects (274). Thus PAR4 AP, but
not PAR1 AP, stimulates release of TNF-␣ from brain
microglia, an effect that is PAR4 specific because it is
prevented by downregulation of PAR4 with antisense.
PAR4 AP also activates the inflammatory NF␬B pathway
603
604
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
suggest that proteases that activate PAR2 can signal to the
peripheral projections of primary spinal afferent neurons
to trigger release of neuropeptides from their peripheral
and central projections; peripheral release causes neurogenic inflammation, and central release induces thermal
hyperalgesia. The identity of the PAR2 activating proteases that mediate these effects remains to be determined, but one possibility is tryptase from mast cells.
Mast cells containing tryptase are in close proximity to
sensory nerve endings in uninflamed and inflamed tissues
(267). Indeed, PAR2-deficient mice show diminished thermal hyperalgesia after intraplantar injection of compound
48/80 to degranulate mast cells, suggesting a role for PAR2
in mast cell-dependent hyperalgesia (296). PAR2 activators also induce pain in the pancreas by a neurogenic
mechanism that involves activation of pancreatic sensory
nerves (124). This observation is relevant to the painful
condition of pancreatitis, where trypsinogen is prematurely activated within the pancreas and could cause hyperalgesia by activating PAR2 on pancreatic neurons.
The molecular mechanism of PAR2-mediated inflammation and hyperalgesia is unknown. However, it is likely
to involve signaling events that regulate activity or expression of ion channels. One candidate is vanilloid receptor-1
(VR1), a member of the transient receptor potential (TRP)
family of channels. TRPV1 is a nonselective cation channel that is activated by protons, elevated temperature, and
certain lipids, as well as by exogenous vanilloids such as
capsaicin (37, 141). TRPV1 plays an important role in
inflammatory pain and thermal hyperalgesia, and several
Physiol Rev • VOL
inflammatory agents transactivate this channel. The
TRPV1 antagonist capsezapine antagonizes PAR2-mediated thermal hyperalgesia (152). Moreover, activation of
PAR2 enhances TRPV1-induced Ca2⫹ signaling and TRPV1
currents in cell lines and neurons by a PKC-dependent
mechanism (Bunnett, unpublished observation) and also
potentiates capsaicin-induced release of CGRP (124).
These results suggest that activated PAR2 couples PKC,
which may phosphorylate and activate TRPV1, thereby
inducing hyperalgesia.
In a similar manner, activation of PAR1 promotes
extravasation of plasma proteins in many tissues by a
mechanism that depends on the release of SP from the
peripheral projections of afferent neurons (87). However,
in contrast to PAR2, PAR1 activation causes somatic analgesia, by mechanisms that remain to be defined (11,
151).
C. Protease Signaling in the
Gastrointestinal System
All PARs are expressed in the gastrointestinal tract,
although most studies have focused on PAR2, which appears to be of particular importance in the gut. The role of
PARs in the gastrointestinal tract has been recently reviewed (295).
1. Proteases
Of all organ systems, the gastrointestinal tract is
exposed to the widest array of proteases under physio-
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
FIG. 8. Protease signal to primary
spinal afferent neurons. 1) Proteases
from mast cells (tryptase), epithelial cells
(trypsin), and the circulation (FVIIa,
FXa) can cleave PAR2 at the peripheral
projections of primary spinal efferent
neurons. 2) Cleavage releases calcitonin
gene-related peptide (CGRP), substance
P (SP), and neurokinin A (NKA). 3)
CGRP interacts with the type 1 CGRP
receptor [comprised of calcitonin receptor like receptor and receptor activity
modifying protein (CRLR/RAMP1)] on arterioles to induce dilation and hyperemia. 4) SP/NKA interact with the neurokinin 1 receptor on endothelial cells of
postcapillary venules to induce gap formation, plasma extravasation, and granulocyte infiltration. 5) PAR2 sensitizes
TRPV1. 6) SP release within the cord
induces thermal hyperalgesia. [Adapted
from Steinhoff et al. (270).]
PROTEASE-ACTIVATED RECEPTORS
2. PARs
PAR1 is present in endothelial cells of the lamina
propria of the small intestine, in intestinal epithelial cells,
smooth muscle cells, and in neurons of the enteric nervous system (24, 67). PAR2 mRNA is strongly expressed in
the small intestine, colon, liver, and pancreas, and more
weakly detected in the stomach (24, 213, 214). Immunoreactive PAR2 is localized to the apical and basolateral
membrane of enterocytes, in smooth muscle cells, on
peripheral terminals of mesenteric afferent nerves, on
neurons of the myenteric and submucosal nerve plexuses,
to endothelial cells and vascular smooth muscle, and to
certain immune cells (neutrophils, mast cells, lymphocytes) (40, 66, 67, 73, 79, 167, 235). PAR3 mRNA is expressed in stomach and small intestine, although the cell
types remain to be defined (135). PAR4 is highly expressed in the pancreas, small intestine, and colon, although it is yet to be localized at the cellular level (143,
317).
3. Ion transport
Activation of PAR1 and PAR2 stimulates Cl⫺ secretion from the intestinal mucosa. The functional implications of this regulation are twofold. First, stimulation of
Cl⫺ secretion would enhance fluid secretion, resulting in
diarrhea that is symptomatic of intestinal inflammation.
Second, enhanced secretion during intestinal inflammation may serve a protective function by promoting the
excretion of toxins.
Physiol Rev • VOL
The effects of PAR activation on electrolyte secretion
by cultured enterocytes and sheets of mucosa have been
examined using Ussing chambers, allowing application of
agonists to apical (lumen) and basolateral (serosal) surface and measurement of short-circuit current as an index
of ion secretion. The role of PAR1 in electrolyte secretion
has been examined using SCBN cells, a nontransformed
epithelial cell line from the crypts of the human small
intestine (30, 31). Application of thrombin or a selective
PAR1 AP to the basolateral surface of these cells stimulates an increase in short-circuit current that is attributable to Cl⫺ secretion by a Ca2⫹-dependent mechanism.
Although SCBN cells also express the cystic fibrosis transmembrane conductance regulator (CFTR), activation of
PAR1 does not raise intracellular cAMP, and thus PAR1 is
unlikely to regulate the CFTR. PAR1 agonists induce activation of Src, transactivation of the EGF receptor, activation of ERK1/2, phosphorylation of cytoplasmic phospholipase A2, and stimulation of cyclooxygenase-1 and -2,
which induce Cl⫺ secretion.
PAR2 signaling in enterocytes has been investigated
using hBRIE cells, a nontransformed epithelial cell line
from the rat small intestine that retains many characteristics of enterocytes. Trypsin and PAR2 AP stimulate generation of InsP3, mobilization of Ca2⫹, and release of
arachidonic acid and prostaglandins E2 and F1␣ (167). The
generation of prostaglandins is also stimulated by the
application of physiological concentrations of trypsin to
the apical membrane of everted sacs of rat jejunum, suggesting expression of PAR2 at the apical membrane. Prostaglandins released from enterocytes may have a protective function and could also regulate fluid and electrolyte
secretion. Application of PAR2 AP and activating proteases to the serosal surface of segments of intestine from
rat (300), mouse (73), and pig (109) induces an increase in
short-circuit current due to stimulation of Cl⫺ secretion,
but the mechanisms vary between species and perhaps
between different regions of the intestine. In the rat jejunum, Cl⫺ secretion depends on generation of prostaglandins (300). However, in view of the very different potencies of PAR2 agonists in this preparation and a PAR2transfected cell line, Cl⫺ secretion may be due to the
activation of a variant of PAR2. In the pig ileum, PAR2stimulated Cl⫺ secretion depends both on eicosanoids
and on submucosal neurons, since the stimulation is
strongly suppressed by the neuronal toxin saxitoxin
(109). Thus application of PAR2 agonists to the serosa
may activate submucosal neurons, which then release
stimulants of Cl⫺ secretion. In the distal colon of the
mouse, application of PAR2 agonists to the serosal surface stimulates Cl⫺ and K⫹ secretion and inhibits amilioride-sensitive Na⫹ absorption (73). These effects are mediated in part by the enteric nervous system.
PAR2 is also expressed by epithelial cells of the pancreatic duct, where it regulates ion transport. Trypsin and
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
logical circumstances and during diseases. Digestive proteases from glands in the stomach and intestine as well as
from extrinsic sources such as the pancreas bathe the
lumen of the gastrointestinal tract during and after feeding. Thus physiological concentrations of trypsin in the
lumen of the small intestine can signal to enterocytes by
cleaving PAR2 at the apical membrane (167). The intestinal lumen is also replete with bacteria and thus exposed
to bacterial proteases. Under normal circumstances, the
tight junctions of epithelial cells act as a barrier to prevent
the ingress of proteases from the lumen into tissues.
However, during stress and inflammation, this barrier is
disrupted and proteases from the lumen may penetrate
the mucosa and muscle layers of the intestine (318).
During inflammation, coagulation proteases and proteases released from inflammatory cells such as mast cells
and neutrophils bathe many cell types and can also be
detected in the intestinal lumen. For example, high levels
of tryptase and trypsin are found in the lumen of the
inflamed intestine (32, 232, 250, 258). Many of these enzymes that are normally present in the intestine or that
are generated or released during inflammation activate
PARs on gastrointestinal cells to influence motility, secretion, and neurotransmission.
605
606
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
PAR2 AP stimulate short-circuit current when added to
the basolateral but not apical surface of monolayers of
duct cells from the canine pancreas by stimulating Ca2⫹activated Cl⫺ channels and K⫹ channels (205). In this
manner PAR2 may play a protective role in pancreatitis,
when there is premature activation of trypsinogen and
secretion of trypsin into the interstitial fluid. Activation of
PAR2 on the basolateral membrane of duct cells could
increase ductal secretion and promote clearance of toxins
and debris.
4. Motility
5. Exocrine secretion
PAR2 regulates pancreatic, salivary, and gastric secretions. PAR2 activation stimulates the release of amylase from isolated rat pancreatic acini (24). Intravenous
injection of PAR2 AP in rats and mice also stimulates
pancreatic amylase secretion, and the effect in mice is
mediated by NO (158). PAR2 is also present in acini of the
salivary gland, and PAR2 APs stimulate release of amylase
and mucus in rats and mice (158). In mice, the effect on
amylase secretion is partially mediated by NO, and the
Physiol Rev • VOL
6. Neurotransmission
The enteric nervous system comprises a series of
ganglionated nerve plexuses that comprise the intrinsic
nervous system of the gut. These ganglia include primary
afferent neurons, interneurons, and motor neurons, allowing local, reflex regulation of all aspects of gastrointestinal function, including control of motility, exocrine and
endocrine secretion, and fluid and electrolyte transport.
The expression of PARs in neurons of the myenteric
plexus suggests a role for proteases in control of motility.
Indeed, PAR1, PAR2, and PAR4 are expressed by a substantial proportion of myenteric neurons in the guinea pig
intestine (⬎50% for PAR1, PAR2) (67, 104, 176). Thrombin,
trypsin, and tryptase induce a slow and prolonged depolarization in myenteric neurons that is often accompanied
by increased excitability (104, 176). Selective APs for
PAR1, PAR2 and PAR4 similarly excite these neurons.
Neurons of the enteric nervous system can be classified
on the basis of their electrophysiological behavior. Two
prominent types of neurons are AH, which are the intrinsic primary afferent neurons, and S neurons, which are
motor neurons and interneurons. Approximately 50 – 60%
of AH neurons and slightly fewer S neurons respond to
PAR2 agonists. The PAR2-responsive S neurons also express NO synthase and project axons in an anal direction.
These are inhibitory motor neurons, and activation would
be expected to suppress motility. PAR2 is also expressed
by neurons of the submucosal nerve plexus of the guinea
pig small intestine, some of which also express vasoactive
intestinal polypeptide and are thus secretomotor neurons
that play an important role in the control of fluid and
electrolyte transport (235). Trypsin, tryptase, and PAR2
AP evoke transient depolarization of submucosal neurons
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
Activation of PAR1 and PAR2 stimulates contraction
of gastric longitudinal muscle from rats and guinea pigs,
which is blocked by the cyclooxygenase inhibitor indomethacin and the tyrosine kinase inhibitor genistein (3,
247). In the mouse gastric fundus, activation of PAR1 and
PAR2 has a biphasic effect, causing relaxation and then
contraction (59). The relaxation is blocked by apamin or
ryanodine, indicating that it is mediated by ryanodinesensitive and -insensitive activation of small-conductance
Ca2⫹-activated K⫹ channels. In the longitudinal muscle of
the rat colon, PAR2 APs inhibit spontaneous contractions
(66). PAR1 and PAR2 APs also inhibit contraction of colonic circular muscle, which is prevented by apamin
(198). Some muscles express more than one receptor for
the same protease, and the net effect on motility depends
on which receptor is activated. In the rat esophagus, PAR1
APs stimulate contraction, but PAR4 APs have little effect
(154). However, in carbachol-precontracted tissues, PAR1
agonists cause additional contraction, whereas PAR4 agonists induce a strong relaxation. These results suggest
that thrombin could have a dual effect in the esophagus,
causing contraction by PAR1 and relaxation by PAR4. The
dominant effect would depend on the concentration of
thrombin: low concentrations preferentially activate
PAR1, but high concentrations of thrombin are needed to
activate PAR4. The intraperitoneal administration of selective APs of PAR1 and PAR2 to mice accelerates intestinal transit (155). These effects are abolished by verapamil, an inhibitor of the L-type Ca2⫹ channel, but are
potentiated by apamin. Thus the net effects of PAR1 and
PAR2 activators in vivo are to promote intestinal transit.
stimulation of mucin secretion in rats depends on activation of tyrosine kinase (157). Activation of PAR2 stimulates secretion of gastric, but not duodenal, mucus (153).
In this manner, PAR2 APs protect the gastric mucosa from
injury induced by acid-ethanol or by indomethacin. PAR2induced mucus secretion and gastric protection are suppressed by antagonists of the CGRP and neurokinin-2
receptors and by the TRPV1 antagonist capsezapine, and
thus depend on the release of neuropeptides from sensory
nerves (152). In contrast, PAR2-mediated salivary secretion is independent of capsaicin-sensitive sensory neurons (156), an observation that illustrates the complexity
of the PAR2 signaling system, which varies between tissues. PAR2 is also expressed by the pepsin-secreting chief
cells of the stomach, and PAR2 agonists stimulate pepsin
secretion in rats (160). Conversely, activation of PAR2
suppresses gastric acid secretion in intact animals (211).
Gastric epithelial cells also express PAR1, and PAR1 AP
stimulates secretion of mucus and prostaglandin E2,
which may be protective (279).
PROTEASE-ACTIVATED RECEPTORS
7. Intestinal inflammation
Since intestinal inflammation is associated with the
generation and release of proteases that are potential
activators of PARs, there has been considerable interest
in the role of PARs in inflammatory diseases of the intestine. Administration of the trypsin, tryptase, and PAR2 AP
into the lumen of the mouse colon provokes a generalized
inflammatory reaction in wild-type but not PAR2-deficient
mice (40, 41). PAR2-induced inflammation is suppressed
by inhibition of NO synthase, by ablation of sensory
nerves, and by antagonism of the CGRP type 1 receptor
and the neurokinin-1 receptor. Thus proteases induce
inflammation in part by a neurogenic mechanism that
involves the release of peptides from sensory nerves.
Intracolonic PAR2 activators also disrupt the intestinal
barrier integrity as observed by the increased paracellular
permeability to bacteria. This disruption of tight junctions
depends on activation of myosin light-chain kinase (41).
Given the large presence and increased activity in the
intestinal lumen of trypsin and tryptase, which can activate PAR2, these data have important implications for the
pathophysiology of inflammatory bowel diseases. However, PAR2 may also serve a protective function in the
intestine. Activation of PAR2 induces release of prostaglandins from enterocytes (167) and stimulates gastric
mucus secretion (153), both of which are protective. Administration of PAR2 AP in mice protects against inflammation induced by the rectal administration of dinitrobenPhysiol Rev • VOL
zene sulfonic acid, a widely used model of hapten-induced
colitis (103). This protective effect is also neurogenic,
since it is prevented by ablation of sensory nerves and
antagonism of the CGRP 1 receptor.
D. Protease Signaling in the Airway
The potential involvement of PARs in regulation of
the airways and in pulmonary disease has been recently
reviewed (58, 127).
1. Proteases
During inflammation and trauma, proteases from the
coagulation cascade and from inflammatory cells could
regulate pulmonary cells by activating PARs. Indeed,
there are elevated levels of coagulation proteases (134)
and tryptase (139) in bronchoalveolar lavage from patients with chronic inflammatory diseases. Proteases that
activate PARs may also arise from cells in the airways.
Immunoreactive trypsinogen colocalizes with PAR2 in
Clara cells of the epithelium of human bronchioles,
whereas in larger bronchioles trypsinogen is expressed by
cells just above the basal layer (56, 58). In addition, a
trypsinlike enzyme can be detected by zymography in the
bronchoalveolar lavage collected from mice treated intranasally with the bacterial endotoxin lipopolysaccharide.
Intriguingly, certain nonmammalian proteases such as
dust mite allergens can also signal to cells in the airway by
cleaving PARs, which may have implications for allergic
inflammation (273).
2. PARs
The PARs are expressed by many cell types in the
airway. There is weak staining for immunoreactive PAR1
in ciliated and basal epithelial cells and in smooth muscle
cells of rat trachea and bronchi (49, 127). In rat, mouse,
guinea pig, and human, immunoreactive PAR2 is expressed by ciliated and nonciliated epithelial cells, especially at the apical membrane, as well as in glands, airway
smooth muscle, and endothelial cells and vascular
smooth muscle cells (49, 56, 58, 79, 192, 237, 255).
3. Airway resistance
Given the importance of increased airway resistance
in asthma, there is great interest in understanding the
regulation of contraction of airway smooth muscle. There
is evidence that PAR2 activators induce both bronchodilatation, which would be protective, and detrimental
bronchoconstriction.
The first evidence in favor of a protective role came
from the finding that activation of PAR2 induces relaxation of isolated bronchi from several species, including
mouse, rat, guinea pig, and human, precontracted with
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
followed by a prolonged hyperexcitability that can last for
several hours. This remarkable long-term hyperexcitability mimics that observed after degranulation of mast cells
in the presence of antagonists of established excitatory
mast cell mediators (histamine, serotonin, prostaglandins). Mast cell degranulation results in release of proteases, which desensitizes neurons to other activators of
PAR2. Thus PAR2 AP and proteases from degranulated
mast cells induce long-term excitability of submucosal
neurons, which could result in enhanced fluid and electrolyte secretion from the mucosa (73, 109). PAR2 is also
expressed by extrinsic neurons that innervate the intestine. Thus the administration of PAR2 activators into the
lumen of the rat intestine induces a hyperalgesia to distension of the intestine with a balloon, consistent with a
visceral hyperalgesia (60). However, in contrast to the
somatic hyperalgesia that depends on the spinal release of
SP, PAR2-dependent visceral hyperalgesia involves peripheral release of SP. The widespread expression of
PARs in the intrinsic and extrinsic nervous systems of the
gut may be of considerable interest during inflammation,
where proteases from the circulation, inflammatory cells,
and the intestinal lumen could alter neuronal excitability
and induce alterations in motility, secretion, and pain
perception.
607
608
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
4. Electrolyte transport
PAR2 also regulates ion transport in the airway epithelium (78, 171). PAR2 activators, added to the basolateral surface of bronchial epithelial cells and airway tissue,
cause a transient Ca2⫹-dependent increase in short-circuit
current, followed by a sustained inhibition of amilioridePhysiol Rev • VOL
sensitive current. This mechanism may control fluid volume and composition at the airway surface.
5. Inflammation and fibrosis
There has been considerable interest in the role of
tryptase in airway inflammation and asthma. Hyperplasia
of smooth muscle contributes to remodeling of the airways that accompanies asthma, and tryptase induces proliferation of airway smooth muscle cells, fibroblasts, and
epithelial cells (29, 33, 244). Tryptase also induces hyperresponsiveness in the airway (17), and when injected into
the skin tryptase causes plasma extravasation and neutrophil infiltration (119, 120). In a sheep model, inhaled
tryptase causes bronchial hyperresponsiveness, whereas
a tryptase inhibitor reduces early and late phase bronchoconstriction and hyperresponsiveness to allergen challenge (55). Although it is not established that these effects
of tryptase are mediated by PAR2, PAR2 AP does have
similar effects. For example, PAR2 AP also induces proliferation of airway smooth muscle cells and lung fibroblasts, and thus PAR2 may mediate the proliferation effects of tryptase on lung fibroblasts and myocytes (2, 18).
PAR2 may have both proinflammatory and protective
effects in the airway. Airway epithelial cells express
PAR1, PAR2, and PAR4, and activation of these receptors
triggers the release of cytokines such as interleukin-6,
interleukin-8, and prostaglandin E2 that can regulate inflammation (12). Observations on PAR2 knockout mice
and animals overexpressing human PAR2 suggest that this
receptor contributes to allergic inflammation (255). In an
ovalbumin-induced allergic inflammation of the airway,
both the infiltration of eosinophils into the airway lumen
and methacholine-induced airway hyperreactivity are
markedly diminished in PAR2-deficient animals, and exacerbated in mice overexpressing PAR2. PAR2 deletion
also reduces IgE levels to ovalbumin sensitization by
fourfold compared with wild-type animals. These results,
together with the observation of upregulation of PAR2 in
airway epithelium of asthmatics (164), suggest that PAR2
contributes to the development of immunity and to allergic inflammation of the airway. However, PAR2 can also
be protective against airway inflammation. Intranasal delivery of PAR2 AP in mice does not cause any inflammation in the airways but inhibits the marked immune cell
inflammatory response, characterized by a recruitment of
polymorphonuclear leukocytes, triggered by administration of lipopolysaccharide (195).
Although less is known about PAR1 in the airway,
there are reports of a role for thrombin and PAR1 in
fibrotic diseases. Thrombin and PAR1 AP are mitogenic
for human airway smooth muscle cells (280). Thrombin
also stimulates the release of the proinflammatory and
fibrogenic cytokine granulocyte-macrophage colony-stimulating factor (GM-CSF) from airway smooth muscle
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
carbachol (49, 56, 58, 172). This relaxation requires the
presence of the epithelium and is mediated by prostanoids but not by NO. Surprisingly, in the rat, PAR2 AP but
not trypsin causes relaxation, possibly due to the existence of trypsin inhibitors in that tissue (49). PAR2-induced relaxation of the isolated mouse trachea is unaffected by acute damage to the epithelium caused either by
mechanical rubbing or infection with influenza A virus
(173). However, the effect of PAR2 activation in the airway may depend on the tissue chosen for study and the
species. For example, whereas PAR2 activation causes
relaxation of isolated guinea pig trachea and main bronchi by a mechanism that is dependent on the presence of
the epithelium and that is mediated by NO and prostanoids, in smaller intrapulmonary bronchi the sole effect of
trypsin and PAR2 AP is to trigger smooth muscle contraction (237). PAR2 activators also stimulate contraction of
lobar or segmental bronchial rings from humans that is
mediated in part by a direct activation of PAR2 on smooth
muscle cells (256). Since smaller bronchioles are the principal site of airway resistance, contraction would be expected to increase total airway resistance, which would
be detrimental. In addition to the aforementioned indirect
effects of PAR2, which often require the presence of the
epithelium, PAR2 activators, including tryptase, can directly signal to isolated smooth muscle cells from the
human airway by mobilizing intracellular Ca2⫹ (19, 256).
Studies in intact animals also show a protective or
detrimental effect of PAR2 activation. In support of a
protective role, aerosol administration of PAR2 AP to
intact rats inhibits serotonin-induced bronchoconstriction (56, 58). Similarly, intravenous administration of
PAR2 AP inhibits histamine-induced increase in airway
resistance in guinea pigs by a mechanism independent of
the release of prostaglandins, NO, or the effect of circulating epinephrine (51). In contrast, intravenous and intratracheal administration PAR2 activators causes bronchoconstriction in intact guinea pigs, mediated in part by
prostanoids and involving the release of tachykinins from
sensory nerve endings and activation of neurokinin-1 and
-2 receptors (237). Thus PAR2 may be protective or detrimental in the airways, depending on the experimental
model and species. It will be of great interest to examine
the contribution of proteases and PAR2 to airway disease
by using specific antagonists and genetically modified
animals.
PROTEASE-ACTIVATED RECEPTORS
E. Protease Signaling in the Skin
The role of proteases in cutaneous biology and disease has recently been reviewed (234).
1. Proteases
Cells in the epidermis and dermis could be regulated
by proteases of the coagulation cascade. Inflammation of
the skin is associated with an influx of inflammatory cells
that release proteases that activate PAR2. Thus whereas
in normal skin mast cells are found in the dermis, in
atopic dermatitis and psoriasis there is an influx of
tryptase-containing mast cells in the dermal and epidermal junction, as well as the epidermis (268). In addition,
epidermal cells express proteases, some of which could
activate PARs. Potential activating enzymes include stratum corneum tryptic enzyme and chymotryptic enzyme
(96, 275). However, it remains to be determined if proteases such as these activate PARs. Exogenous proteases
from mites, bacteria, and fungi could also signal to epidermal cells.
PAR2 (126, 268). The expression of PAR2 has been evaluated in some skin diseases (268). Thus PAR2 immunoreactivity is strong in the hyperproliferative granular layer
of skin from patients with lichen planus, and in atopic
dermatitis there is strong staining of the whole epidermis,
as well as in blood vessels and dendritic-like cells.
3. Pigmentation
There is considerable interest in the role of PAR2 in
pigmentation. The interaction between keratinocytes and
neighboring melanocytes is essential for skin pigmentation. Melanocytes produce secretory granules, melanosomes, which produce melanin. Melanin-containing melanosomes are transferred to dendrites of melanocytes and
are then taken up by keratinocytes by phagocytosis, resulting in skin darkening. Although melanocytes do not
express PAR2, PAR2 activators regulate the phagocytosis
of melanosomes by keratinocytes and thereby control
pigmentation (259, 264). Thus trypsin and PAR2 AP stimulate pigmentation of cocultures of keratinocytes and
melanocytes. Remarkably, the topical application of PAR2
AP induces pigmentation of human skin transplanted
onto mice. Conversely, a serine protease inhibitor causes
lightening of the skin (224). The mechanism of these
effects of PAR2 on pigmentation depends on the stimulation of phagocytosis in keratinocytes. Thus activation of
PAR2 increases the ingestion of particles and bacteria by
keratinocytes (264). Phagocytosis by keratinocytes not
only plays a role in pigmentation but also enables “nonimmune” cells such as these to participate in protection of
the skin, with implications for wound healing and inflammation. Of particular interest, the increased phagocytic
response of keratinocytes correlates with increased activity of soluble serine proteases. Serine protease inhibitors
downregulated both constitutive and PAR2 stimulated
phagocytosis, suggesting that these proteases also activate PAR2 and thereby amplify the response. PAR2 is
upregulated in keratinocytes after ultraviolet exposure,
with implications for pigmentation (257).
4. Proliferation and wound healing
2. PARs
Immunoreactive PAR1 is found in the basal and spinous layers of the human epidermis but not in the granular layer, and human basal keratinocytes and a keratinocytes cell line express PAR1 mRNA and protein (76).
PAR2 immunoreactivity is predominantly localized to the
basal and suprabasal spinous layers, although the stratum
corneum is unstained (76, 268). PAR2 is also expressed in
the inner root sheath of hair follicles and in myoepithelial
cells of sweat glands. Keratinocytes in culture also express PAR2 mRNA and protein (251). PAR2 mRNA is also
moderately expressed in cultured human microvasculature endothelial cells, but melanocytes do not express
Physiol Rev • VOL
Several findings suggest that PAR1 could play a role
in wound healing. Activation of PAR1 stimulates proliferation of keratinocytes (5, 251) and fibroblasts (45), and
also induce angiogenesis (285). Plasmin, thrombin, and
PAR1 AP induce expression of cysteine-rich angiogenic
protein Cyr-61, a growth factor-like gene implicated in
angiogenesis and wound healing, in fibroblasts from wildtype but not PAR1-deficient mice (227). Moreover, the
topical application of thrombin and PAR1 AP to incisional
wounds in rats improves wound strength and increases
angiogenesis, thereby promoting wound healing (36). The
role of PAR1 in wound healing has been examined in
PAR1-deficient mice (65). Although fibroblasts from these
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
cells. Although this effect is blocked by pertussis toxin,
and thus likely mediated by a GPCR, it is not mimicked by
activators of PAR1 and PAR4, and may thus involve a
novel receptor. However, PAR1 may have a role in fibrotic
diseases of the lungs. Instillation of bleomycin into the
airway induces pulmonary fibrosis in rats, which is associated with elevated expression of thrombin and PAR1 in
macrophages and fibroblasts (127). Administration of a
thrombin inhibitor UK-156406 attenuates bleomycin-stimulated deposition of collagen in the lung by 40%, suggesting an important role of coagulation proteases in fibrogenesis in the lung. Moreover, thrombin and FXa stimulate expression of connective tissue growth factor from
fibroblasts by activation PAR1, which may contribute to
normal repair and fibrosis (42). Inhibitors of coagulation
proteases and antagonists of PAR1 could be novel therapies for pulmonary fibrosis.
609
610
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
animals do not respond to thrombin, the time to closure
of open skin wounds and the incision strength of healed
wounds is unaffected by deletion of PAR1. The discrepancy between this observation and the finding that PAR1
AP promotes healing in rats is difficult to explain, but
could be related to differences in species and the wound
model chosen.
5. Inflammation
VI. PROTEASE-ACTIVATED RECEPTORS:
MECHANISMS OF DISEASE AND TARGETS
FOR THERAPY
Proteases and PARs make important contributions to
disease and are targets for therapies. First, PARs regulate
many biological processes that are critical in disease,
including coagulation, proliferation and survival, inflammation, neurotransmission, and pain. Second, proteases
that activate PARs are generated during diseases, for
Physiol Rev • VOL
A. Cardiovascular Disease
The inappropriate aggregation of platelets makes an
important contribution to occlusive vascular disorders
such as stroke, angina, and myocardial infarction, where
accumulation of atherosclerotic plaques promotes platelet-mediated thrombus formation. Thrombin is a major
mediator of platelet aggregation and fibrin deposition, and
thrombin inhibitors are useful antithrombotic agents. One
drawback of such inhibitors is their disruption of the
normal hemostatic mechanism, and a selective suppression of the effects of thrombin on platelets may be advantageous. A difficulty of studying the role of PARs in thrombus formation is that platelets express several thrombin
receptors, which vary among species. However, despite
the redundancy and interspecies differences, recent observations in cynomologus monkeys, whose platelets, like
those of humans, express PAR1 and PAR4, suggest that
PAR1 antagonists may be useful therapeutic agents in
humans (88). Thus the PAR1 antagonist RWJ-58259 markedly inhibits thrombus formation and vessel occlusion in
a model of electrolytic injury of the carotid artery in
cynomologus monkeys. Despite the fact that platelets
from these animals possess the dual PAR1/PAR4 system,
antagonism of a single receptor is effective, raising the
possibility that antagonism of a PAR1 may be of use for
treatment of thrombosis disorders in humans. In guinea
pigs, whose platelets express three thrombin receptors
(PAR1, PAR3, and PAR4), RWJ-58259 is only marginally
effective in arterial thrombosis, which emphasizes the
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
Activation of PAR1 and PAR2 induces expression of
interleukin-6 and -8 as well as GM-CSF in cultured human
keratinocytes, indicating that serine proteases modulate
cytokine expression in the epidermis (126, 307). GM-CSF
mediates Langerhans cell maturation, activates tissue
macrophages, and stimulates proliferation of keratinocytes. PAR2 agonists also activate stress-activated protein
kinases (c-Jun amino-terminal kinase or JNK and p38) and
the NF␬B pathways in the keratinocyte cell line
NCTC2544 (145). The NF␬B family of transcription factors is implicated in regulation of a number of proinflammatory genes in these cells.
The proinflammatory effects of PAR2 have also been
examined in the intact animal. PAR2-deficient mice exhibit diminished ear swelling and infiltration of inflammatory cells in a model of allergic dermatitis, suggesting that
PAR2 mediates allergic dermatitis (159). PAR2 activators
also signal to primary spinal afferent neurons that innervate the skin, resulting in the release of the neuropeptides
SP and CGRP, which together induce neurogenic inflammation (270) (Fig. 8). This signaling also triggers a prolonged mechanical and thermal somatic hyperalgesia,
which is due to the release of SP from central projections
of these neurons within the dorsal horn of the spinal cord
(296). Protease signaling to sensory nerves in the skin has
recently been investigated in humans (234). Atopic dermatitis is associated with an elevated release of tryptase
in the skin and an upregulation of PAR2 on nerve cells.
Cutaneous administration of PAR2 AP to these patients
results in prolonged itch, even in the presence of antihistamine drugs (269). These results suggest that tryptase
could act on PAR2-expressing sensory nerves in human
skin to produce itch, which may explain the insensitivity
of these patients to antihistamines.
example, in trauma, hemostasis, inflammation, and tumor
formation. Third, the expression and function of PARs are
altered in disease. Thus PAR2 is upregulated during inflammation (216), and a polymorphic variant of PAR2 has
been identified with markedly abnormal sensitivity to
PAR agonists (61). Finally, studies of genetically modified
animals and using PAR antagonists and protease inhibitors suggest a role of PARs in models of diseases.
However, there are formidable obstacles toward understanding of the role of proteases and PARs in disease.
A major problem is the lack of widely available antagonists; although PAR-deficient mice are available, mice are
not always the optimal species to model human diseases.
Another problem is the apparent redundancy of the PAR
system. Cells express several PARs for the same protease,
and several proteases can activate the same receptor.
Effective blockade of protease signaling may require antagonism of several PARs or proteases. Finally, the roles
of PARs have been specifically studied in few animal
models of disease. Moreover, little is known about the
expression of proteases and PARs in diseased human
tissues. With these caveats in mind, the role of PARs will
be briefly discussed in several diseases.
PROTEASE-ACTIVATED RECEPTORS
B. Inflammatory Disease
There is considerable interest in understanding the
role of PARs, in particular PAR2, in inflammatory diseases, and this topic has been thoroughly reviewed (54,
57, 69). Observations on PAR2-deficient mice suggest a
role for this receptor in inflammation of the airway, joints,
and kidney. In allergic inflammation of the airway induced
by immunization and challenge with ovalbumin, there is
diminished infiltration of eosinophils and reduced hyperreactivity to methacholine in PAR2-deficient animals,
whereas these responses are exacerbated in mice overexpressing PAR2 (255). Rheumatoid arthritis is a chronic
inflammatory condition that is associated with inflammation of the synovium and hyperplasia. There is also elevated expression of TF and thrombin in the synovium and
the deposition of fibrin in the inflamed joint, indicating
activation of the coagulation system. Immunization and
subsequent challenge of mice with chicken collagen results in arthritis, and administration of the thrombin inhibitor hirudin reduces the severity of the inflammation,
assessed by clinical scoring and measurement of expression of proinflammatory cytokines, providing direct evidence for the involvement of thrombin in the inflammation (186). Evidence for a role of PAR2 in arthritis has
been provided from observations of PAR2-deficient animals, which are protected against arthritis induced by
Physiol Rev • VOL
intra-articular and periarticular injection of adjuvant
(101). This inflammation is accompanied by an upregulation of PAR2 expression, which is normally confined to
the vasculature of the joints. The protease that activates
PAR2 in the inflamed joint remains to be determined,
although it is tempting to speculate that FVIIa and FXa
could be involved in light of the established activation of
the coagulation cascade in arthritis. Proteases and PARs
may also contribute to renal inflammation. Cresentic glomerulonephritis is a severe and progressive form of renal
inflammation that is a common cause of end-stage kidney
disease. It is associated with fibrin deposition, suggesting
activation of the coagulation proteases. Immunization
with sheep globulin followed by challenge with sheepanti-mouse glomerular basement membrane induces glomerulonephritis in mice that is characterized by fibrin
deposition, infiltration of leukocytes, and reduced glomerular filtration (75). This inflammation is markedly diminished by treatment with the thrombin inhibitor hirudin,
and inflammation is reduced in PAR1-deficient mice, indicating involvement of coagulation proteases and PAR1.
The observation that inflammation is diminished in
PAR-deficient animals suggests that PAR antagonists and
protease inhibitors may be useful anti-inflammatory
agents. Indeed, tryptase inhibitors have been used to treat
asthma and inflammatory bowel disease in humans (170,
284). However, it must be emphasized that PARs can
exert anti-inflammatory and protective effects in certain
conditions. Thus administration of recombinant APC
blocks apoptosis in human brain endothelial cells and is
neuroprotective in an animal model of ischemic brain
injury (46). Moreover, APC reduces mortality of patients
with sepsis (20). These protective actions of APC on
endothelial cells are dependent on PAR1 (46, 238). In
addition, activation of PAR2 induces bronchodilatation
(56), inhibits lipopolysaccharide-induced pulmonary neutrophilia (195), diminishes formation of gastric ulcers
(153) and experimental colitis (103), and protects against
ischemia-perfusion injury in the heart (203). Thus agonists of PAR1 and PAR2 could be useful therapies for
certain conditions.
C. Cancer
The microenvironment of tumors is replete with proteases that can activate PARs, and tumor cells themselves
express PARs. Malignant cells secrete thrombin and trypsin, which can affect proliferation and mediate metastatic
processes such as cellular invasion, extracellular matrix
degradation, angiogenesis, and tissue remodeling.
PAR1 and PAR2 are expressed by a wide range of
tumor cells (24, 80, 82, 208, 316). In breast tumor tissues,
PAR1 and PAR2 are expressed in the tumor cells, mast
cells, macrophages, endothelial cells, and vascular
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
importance of selecting the appropriate species to model
human disease (8).
PAR antagonists may also be of interest in treating
vascular injury. In addition to its role in thrombus formation, thrombin is also implicated in restenosis of the blood
vessels after vascular injury. Thus a PAR1 blocking antibody or a PAR1 antagonist attenuates thickening of the
neointima of the rat carotid artery rat, initiated by damage
to the endothelium by balloon angioplasty (8). Similarly,
vascular injury is attenuated in PAR1-deficient mice (301).
The protective effect of PAR1 antagonism or deletion
occurs even though rat and mouse platelets do not express PAR1 and thus exhibit normal aggregation in the
absence of functional PAR1. These effects of thrombin on
restenosis are therefore dependent on the proliferative or
proinflammatory effects of PAR1 on vascular smooth
muscle.
PAR2 may contribute to injury that follows ischemia
and reperfusion of tissues (203). Ischemia and reperfusion of the heart induces injury that is characterized by
generalized inflammation and necrosis. PAR2 is upregulated in a model of coronary ischemia and reperfusion
injury in rats, and administration of a PAR2 AP markedly
protects the heart from injury. PAR2 activation also improves efficiency of ischemic preconditioning and reduces cardiac inflammation in the rat heart (204).
611
612
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
VII. CONCLUSIONS AND
FUTURE PERSPECTIVES
There have been substantive advances in our understanding of protease signaling. Four PARs have been
cloned and characterized, and the mechanisms by which
proteases cleave and activate PARs are understood in
considerable detail. A number of mammalian and nonmammalian proteases have been identified that can activate PARs, providing new insights into protease signaling.
The realization that anchoring proteins, including PARs
themselves or other nonreceptor proteins, play a critical
Physiol Rev • VOL
role as cofactors for protease signaling has led to the
identification of new activating enzymes. PARs have been
used as model receptors to investigate mechanisms of
desensitization and downregulation of GPCRs in general.
Tools have been developed to examine receptor functions, including specific agonists and antagonists and both
knockout and transgenic mice. With the use of these
tools, the contribution of PARs to disease has been investigated in experimental animals. Finally, protease inhibitors, for example, tryptase inhibitors, and proteases themselves, such as APC, have been used to successfully treat
human diseases. The effectiveness of these therapies may
well be attributable to effects on signaling by PARs, raising the possibility that antagonists and agonists of PARs
may be useful drugs.
Despite this impressive progress, there is much to
learn about proteases and their receptors. First, the proteases that activate PARs are not fully characterized.
Given the extremely widespread distribution of PARs, it is
almost certain that new activating enzymes await discovery. This omission is particularly true for PAR2; although
trypsin is a very potent agonist of this receptor, almost
nothing is known about the function and regulation of
extrapancreatic trypsins, which are likely to activate
PAR2 in certain tissues. Second, there is good evidence
for the existence of other PARs. Many proteases have
biological actions that appear to be receptor mediated but
which cannot be explained by the known PARs. Moreover, differences in the rank order of potencies of PAR
agonists in bioassays point to the existence of receptor
subtypes. Whether these novel receptors are PAR-like
GPCRs or different types of receptor remains to be determined. Third, the mechanisms of PAR signaling are not
fully explored. Thus the process by which proteolytic
cleavage of an extracellular portion of the receptor induces the interaction of intracellular domains with heterotrimeric G proteins is not understood. Furthermore,
the molecular events that terminate this irreversible
mechanism of activation are not completely defined. For
example, little is known about the postendocytic sorting
mechanisms that target PARs for degradation. Fourth,
and most important, from the standpoint of physiology
and medicine, the optimal tools to study the function of
PARs, receptor antagonists, are generally lacking. Although antagonists have been reported for certain PARs,
they are not widely available. Finally, information about
the role of proteases and PARs in many experimental
systems is simply lacking. It will be of great interest to
investigate the contributions of proteases, protease inhibitors, cofactors, and PARs in experimental models of
human disease using genetically modified animals and
available pharmacological reagents.
We thank members of our laboratory for valuable discussion and advice and for critically reading this manuscript.
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
smooth muscle cells of the metastatic tumor microenvironment (80). In particular, there is an upregulation of
PAR1 and PAR2 in proliferating stromal fibroblasts surrounding the carcinoma cells. In pulmonary tumor alveolar walls, the expression of PAR1 and PAR2 mRNA is
increased by 10- and 16-fold, respectively, compared with
normal alveolar tissues (140). Expression of trypsin is
also detected in this tumor tissue.
The metastasis of tumor cells requires cell detachment from the expanding tumor mass, degradation of the
stromal tissue and basement membrane, and increased
cell motility. There is considerable evidence that proteases and PAR1 may contribute to these processes. The
level of expression of PAR1 on tumor cells directly correlates with metastatic potential in both primary breast
carcinoma and established cancer cell lines (97). Introduction of PAR1 antisense cDNA inhibits the invasion of
metastatic breast carcinoma cells in culture through a
reconstituted basement membrane, indicating an important role for PAR1 in tumor cell migration. In support of
this observation, thrombin and PAR1 AP promote the
invasion of a highly aggressive breast cancer cell line
expressing PAR1 in an in vitro assay (122). Exposure of
certain tumor cells to thrombin and PAR1 AP also enhances their adhesion to platelets, fibronectin, and von
Willebrand factor in vitro and promotes pulmonary metastasis when cells are administered to mice in vivo (209,
210). Moreover, overexpression of PAR1 in certain tumor
cells can enhance their metastatic potential in animal
models. In colon cancer cell lines, activation of PAR1
induces a marked mitogenic response, which is dependent on activation of MAP kinase ERK1/2, and also stimulate motility of wounded cells (81). Together, these results suggest that antagonists of PAR1 may be useful
treatments for proliferation and metastases of certain
tumors. However, PAR1 AP has also reported to inhibit
migration and invasion of breast cancer cell lines when
applied as a concentration gradient in the direction of
movement (144). PAR2 may also contribute to tumor formation and metastasis since PAR2 AP stimulates proliferation of colon tumor cell lines (82).
PROTEASE-ACTIVATED RECEPTORS
This review is not exhaustive; some papers have been
omitted due to space limitations.
Research in the authors’ laboratory in this area is supported by National Institute of Diabetes and Digestive and Kidney Diseases Grants DK-43207, DK-57489, and DK-39957 and by
the RW Johnson Focused Giving Program (to N. W. Bunnett).
Address for reprint requests and other correspondence:
N. W. Bunnett, Room C317, UCSF, 513 Parnassus Ave., San
Francisco, CA 94143-0660 (E-mail: [email protected]).
16.
17.
REFERENCES
Physiol Rev • VOL
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
Wadsworth RM, Gould GW, and Plevin R. Trypsin stimulates
proteinase-activated receptor-2-dependent and -independent activation of mitogen-activated protein kinases. Biochem J 320: 939 –
946, 1996.
Benka ML, Lee M, Wang GR, Buckman S, Burlacu A, Cole L,
DePina A, Dias P, Granger A, Grant B, Hayward-Lester A,
Karki S, Mann S, Marcu O, Nussenzweig A, Piepenhagen P,
Raje M, Roegiers F, Rybak S, Salic A, Smith-Hall J, Waters J,
Yamamoto N, Yanowitz J, Yeow K, Busa WB, and Mendelsohn
ME. The thrombin receptor in human platelets is coupled to a GTP
binding protein of the G alpha q family. FEBS Lett 363: 49 –52, 1995.
Berger P, Compton SJ, Molimard M, Walls AF, N’Guyen C,
Marthan R, and Tunon-De-Lara JM. Mast cell tryptase as a
mediator of hyperresponsiveness in human isolated bronchi. Clin
Exp Allergy 29: 804 – 812, 1999.
Berger P, Perng DW, Thabrew H, Compton SJ, Cairns JA,
McEuen AR, Marthan R, Tunon De Lara JM, and Walls AF.
Tryptase and agonists of PAR-2 induce the proliferation of human
airway smooth muscle cells. J Appl Physiol 91: 1372–1379, 2001.
Berger P, Tunon-De-Lara JM, Savineau JP, and Marthan R.
Selected contribution: tryptase-induced PAR-2-mediated Ca2⫹ signaling in human airway smooth muscle cells. J Appl Physiol 91:
995–1003, 2001.
Bernard GR, Vincent JL, Laterre PF, LaRosa SP, Dhainaut JF,
Lopez-Rodriguez A, Steingrub JS, Garber GE, Helterbrand
JD, Ely EW, and Fisher CJ Jr. Efficacy and safety of recombinant human activated protein C for severe sepsis. N Engl J Med
344: 699 –709, 2001.
Blackhart BD, Emilsson K, Nguyen D, Teng W, Martelli AJ,
Nystedt S, Sundelin J, and Scarborough RM. Ligand crossreactivity within the protease-activated receptor family. J Biol
Chem 271: 16466 –16471, 1996.
Bohm SK, Grady EF, and Bunnett NW. Regulatory mechanisms
that modulate signalling by G-protein-coupled receptors. Biochem
J 322: 1–18, 1997.
Bohm SK, Khitin LM, Grady EF, Aponte G, Payan DG, and
Bunnett NW. Mechanisms of desensitization and resensitization of
proteinase-activated receptor-2. J Biol Chem 271: 22003–22016,
1996.
Bohm SK, Kong W, Bromme D, Smeekens SP, Anderson DC,
Connolly A, Kahn M, Nelken NA, Coughlin SR, Payan DG, and
Bunnett NW. Molecular cloning, expression and potential functions of the human proteinase-activated receptor-2. Biochem J 314:
1009 –1016, 1996.
Bouton MC, Jandrot-Perrus M, Moog S, Cazenave JP, Guillin
MC, and Lanza F. Thrombin interaction with a recombinant Nterminal extracellular domain of the thrombin receptor in an acellular system. Biochem J 305: 635– 641, 1995.
Brass LF, Pizarro S, Ahuja M, Belmonte E, Blanchard N,
Stadel JM, and Hoxie JA. Changes in the structure and function
of the human thrombin receptor during receptor activation, internalization, and recycling. J Biol Chem 269: 2943–2952, 1994.
Bretschneider E, Kaufmann R, Braun M, Nowak G, Glusa E,
and Schror K. Evidence for functionally active protease-activated
receptor-4 (PAR-4) in human vascular smooth muscle cells. Br J
Pharmacol 132: 1441–1446, 2001.
Bretschneider E, Kaufmann R, Braun M, Wittpoth M, Glusa E,
Nowak G, and Schror K. Evidence for proteinase-activated receptor-2 (PAR-2)-mediated mitogenesis in coronary artery smooth
muscle cells. Br J Pharmacol 126: 1735–1740, 1999.
Brown JK, Tyler CL, Jones CA, Ruoss SJ, Hartmann T, and
Caughey GH. 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.
Buresi MC, Buret AG, Hollenberg MD, and MacNaughton WK.
Activation of proteinase-activated receptor 1 stimulates epithelial
chloride secretion through a unique MAP kinase- and cyclo-oxygenase-dependent pathway. FASEB J 16: 1515–1525, 2002.
Buresi MC, Schleihauf E, Vergnolle N, Buret A, Wallace JL,
Hollenberg MD, and MacNaughton WK. Protease-activated receptor-1 stimulates Ca2⫹-dependent Cl⫺ secretion in human intestinal epithelial cells. Am J Physiol Gastrointest Liver Physiol 281:
G323–G332, 2001.
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
1. Abraham LA, Chinni C, Jenkins AL, Lourbakos A, Ally N, Pike
RN, and Mackie EJ. Expression of protease-activated receptor-2
by osteoblasts. Bone 26: 7–14, 2000.
2. Akers IA, Parsons M, Hill MR, Hollenberg MD, Sanjar S,
Laurent GJ, and McAnulty RJ. Mast cell tryptase stimulates
human lung fibroblast proliferation via protease-activated receptor-2. Am J Physiol Lung Cell Mol Physiol 278: L193–L201, 2000.
3. Al-Ani B, Saifeddine M, and Hollenberg MD. Detection of functional receptors for the proteinase-activated-receptor-2-activating
polypeptide, SLIGRL-NH2, in rat vascular and gastric smooth muscle. Can J Phys Pharmacol 73: 1203–1207, 1995.
4. Al-Ani B, Saifeddine M, Kawabata A, and Hollenberg MD.
Proteinase activated receptor 2: role of extracellular loop 2 for
ligand-mediated activation. Br J Pharmacol 128: 1105–1113, 1999.
5. Algermissen B, Sitzmann J, Nurnberg W, Laubscher JC, Henz
BM, and Bauer F. Distribution and potential biologic function of
the thrombin receptor PAR-1 on human keratinocytes. Arch Dermatol Res 292: 488 – 495, 2000.
6. Alm AK, Gagnemo-Persson R, Sorsa T, and Sundelin J. Extrapancreatic trypsin-2 cleaves proteinase-activated receptor-2.
Biochem Biophys Res Commun 275: 77– 83, 2000.
7. Andersen H, Greenberg DL, Fujikawa K, Xu W, Chung DW,
and Davie EW. Protease-activated receptor 1 is the primary mediator of thrombin-stimulated platelet procoagulant activity. Proc
Natl Acad Sci USA 96: 11189 –11193, 1999.
8. Andrade-Gordon P, Derian CK, Maryanoff BE, Zhang HC,
Addo MF, Cheung W, Damiano BP, D’Andrea MR, Darrow AL,
de Garavilla L, Eckardt AJ, Giardino EC, Haertlein BJ, and
McComsey DF. Administration of a potent antagonist of proteaseactivated receptor-1 (PAR-1) attenuates vascular restenosis following balloon angioplasty in rats. J Pharmacol Exp Ther 298: 34 – 42,
2001.
9. Andrade-Gordon P, Maryanoff BE, Derian CK, Zhang HC,
Addo MF, Darrow AL, Eckardt AJ, Hoekstra WJ, McComsey
DF, Oksenberg D, Reynolds EE, Santulli RJ, Scarborough
RM, Smith CE, and White KB. Design, synthesis, and biological
characterization of a peptide-mimetic antagonist for a tetheredligand receptor. Proc Natl Acad Sci USA 96: 12257–12262, 1999.
10. Aragay AM, Collins LR, Post GR, Watson AJ, Feramisco JR,
Brown JH, and Simon MI. G12 requirement for thrombin-stimulated gene expression and DNA synthesis in 1321N1 astrocytoma
cells. J Biol Chem 270: 20073–20077, 1995.
11. Asfaha S, Brussee V, Chapman K, Zochodne DW, and
Vergnolle N. Proteinase-activated receptor-1 agonists attenuate
nociception in response to noxious stimuli. Br J Pharmacol 135:
1101–1106, 2002.
12. Asokananthan N, Graham PT, Fink J, Knight DA, Bakker AJ,
McWilliam AS, Thompson PJ, and Stewart GA. Activation of
protease-activated receptor (PAR)-1, PAR-2, and PAR-4 stimulates
IL-6, IL-8, and prostaglandin E2 release from human respiratory
epithelial cells. J Immunol 168: 3577–3585, 2002.
13. Baffy G, Yang L, Raj S, Manning DR, and Williamson JR. G
protein coupling to the thrombin receptor in Chinese hamster lung
fibroblasts. J Biol Chem 269: 8483– 8487, 1994.
14. Beecher KL, Andersen TT, Fenton JWN, and Festoff BW.
Thrombin receptor peptides induce shape change in neonatal murine astrocytes in culture. J Neurosci Res 37: 108 –115, 1994.
15. Belham CM, Tate RJ, Scott PH, Pemberton AD, Miller HR,
613
614
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
Physiol Rev • VOL
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
bin receptor activating peptide in guinea-pigs in vivo. Br J Pharmacol 126: 478 – 484, 1999.
Cicala C, Morello S, Santagada V, Caliendo G, Sorrentino L,
and Cirino G. Pharmacological dissection of vascular effects
caused by activation of protease-activated receptors 1 and 2 in
anesthetized rats. FASEB J 15: 1433–1435, 2001.
Cicala C, Pinto A, Bucci M, Sorrentino R, Walker B, Harriot P,
Cruchley A, Kapas S, Howells GL, and Cirino G. Proteaseactivated receptor-2 involvement in hypotension in normal and
endotoxemic rats in vivo. Circulation 99: 2590 –2597, 1999.
Cirino G, Bucci M, Cicala C, and Napoli C. Inflammation-coagulation network: are serine protease receptors the knot? Trends
Pharmacol Sci 21: 170 –172, 2000.
Clark JM, Abraham WM, Fishman CE, Forteza R, Ahmed A,
Cortes A, Warne RL, Moore WR, and Tanaka RD. Tryptase
inhibitors block allergen-induced airway and inflammatory responses in allergic sheep. Am J Respir Crit Care Med 152: 2076 –
2083, 1995.
Cocks TM, Fong B, Chow JM, Anderson GP, Frauman AG,
Goldie RG, Henry PJ, Carr MJ, Hamilton JR, and Moffatt JD.
A protective role for protease-activated receptors in the airways.
Nature 398: 156 –160, 1999.
Cocks TM and Moffatt JD. Protease-activated receptors: sentries
for inflammation? Trends Pharmacol Sci 21: 103–108, 2000.
Cocks TM and Moffatt JD. Protease-activated receptor-2 (PAR2)
in the airways. Pulm Pharmacol Ther 14: 183–191, 2001.
Cocks TM, Sozzi V, Moffatt JD, and Selemidis S. Proteaseactivated receptors mediate apamin-sensitive relaxation of mouse
and guinea pig gastrointestinal smooth muscle. Gastroenterology
116: 586 –592, 1999.
Coelho AM, Vergnolle N, Guiard B, Fioramonti J, and Bueno
L. Proteinases and proteinase-activated receptor 2: a possible role
to promote visceral hyperalgesia in rats. Gastroenterology 122:
1035–1047, 2002.
Compton SJ, Cairns JA, Palmer KJ, Al-Ani B, Hollenberg MD,
and Walls AF. A polymorphic protease-activated receptor 2
(PAR2) displaying reduced sensitivity to trypsin and differential
responses to PAR agonists. J Biol Chem 275: 39207–39212, 2000.
Compton SJ, Renaux B, Wijesuriya SJ, and Hollenberg MD.
Glycosylation and the activation of proteinase-activated receptor 2
[PAR(2)] by human mast cell tryptase. Br J Pharmacol 134: 705–
718, 2001.
Compton SJ, Sandhu S, Wijesuriya SJ, and Hollenberg MD.
Glycosylation of human proteinase-activated receptor-2 (hPAR2):
role in cell surface expression and signalling. Biochem J 368:
495–505, 2002.
Connolly AJ, Ishihara H, Kahn ML, Farese RV Jr, and Coughlin SR. Role of the thrombin receptor in development and evidence
for a second receptor. Nature 381: 516 –519, 1996.
Connolly AJ, Suh DY, Hunt TK, and Coughlin SR. Mice lacking
the thrombin receptor, PAR1, have normal skin wound healing.
Am J Pathol 151: 1199 –1204, 1997.
Corvera CU, Dery O, McConalogue K, Bohm SK, Khitin LM,
Caughey GH, Payan DG, and Bunnett NW. Mast cell tryptase
regulates rat colonic myocytes through proteinase-activated receptor 2. J Clin Invest 100: 1383–1393, 1997.
Corvera CU, Dery O, McConalogue K, Gamp P, Thoma M,
Al-Ani B, Caughey GH, Hollenberg MD, and Bunnett NW.
Thrombin and mast cell tryptase regulate guinea-pig myenteric
neurons through proteinase-activated receptors-1 and -2. J Physiol
517: 741–756, 1999.
Coughlin SR. Thrombin signalling and protease-activated receptors. Nature 407: 258 –264, 2000.
Coughlin SR and Camerer E. PARticipation in inflammation.
J Clin Invest 111: 25–27, 2003.
Covic L, Gresser AL, and Kuliopulos A. Biphasic kinetics of
activation and signaling for PAR1 and PAR4 thrombin receptors in
platelets. Biochemistry 39: 5458 –5467, 2000.
Covic L, Gresser AL, Talavera J, Swift S, and Kuliopulos A.
Activation and inhibition of G protein-coupled receptors by cellpenetrating membrane-tethered peptides. Proc Natl Acad Sci USA
99: 643– 648, 2002.
Covic L, Misra M, Badar J, Singh C, and Kuliopulos A. Pepdu-
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
32. Bustos D, Negri G, De Paula JA, Di Carlo M, Yapur V, Facente
A, and De Paula A. Colonic proteinases: increased activity in
patients with ulcerative colitis. Medicina 58: 262–264, 1998.
33. Cairns JA and Walls AF. Mast cell tryptase is a mitogen for
epithelial cells. Stimulation of IL-8 production and intercellular
adhesion molecule-1 expression. J Immunol 156: 275–283, 1996.
34. Camerer E, Huang W, and Coughlin SR. Tissue factor- and
factor X-dependent activation of protease-activated receptor 2 by
factor VIIa. Proc Natl Acad Sci USA 97: 5255–5260, 2000.
35. Camerer E, Kataoka H, Kahn M, Lease K, and Coughlin SR.
Genetic evidence that protease-activated receptors mediate factor
Xa signaling in endothelial cells. J Biol Chem 277: 16081–16087,
2002.
36. Carney DH, Mann R, Redin WR, Pernia SD, Berry D, Heggers
JP, Hayward PG, Robson MC, Christie J, Annable C, Fenton
JW II, and Glenn KC. Enhancement of incisional wound healing
and neovascularization in normal rats by thrombin and synthetic
thrombin receptor-activating peptides. J Clin Invest 89: 1469 –1477,
1992.
37. Caterina MJ, Schumacher MA, Tominaga M, Rosen TA, Levine JD, and Julius D. The capsaicin receptor: a heat-activated
ion channel in the pain pathway. Nature 389: 816 – 824, 1997.
38. Caughey GH. Mast cell chymases and tryptases: phylogeny, family
relations, and biogenesis. In: Mast Cell Proteases in Immunology
and Biology, edited by Caughey GH. New York: Dekker, 1995, p.
305–329.
39. Cavanaugh KP, Gurwitz D, Cunningham DD, and Bradshaw
RA. Reciprocal modulation of astrocyte stellation by thrombin and
protease nexin-1. J Neurochem 54: 1735–1743, 1990.
40. Cenac N, Coelho AM, Nguyen C, Compton S, Andrade-Gordon
P, MacNaughton WK, Wallace JL, Hollenberg MD, Bunnett
NW, Garcia-Villar R, Bueno L, and Vergnolle N. Induction of
intestinal inflammation in mouse by activation of proteinase-activated receptor-2. Am J Pathol 161: 1903–1915, 2002.
41. Cenac N, Garcia-Villar R, Ferrier L, Larauche M, Vergnolle N,
Bunnett NW, Coelho AM, Fioramonti J, and Bueno L. Proteinase-activated receptor-2-induced colonic inflammation in mice:
possible involvement of afferent neurons, nitric oxide, and paracellular permeability. J Immunol 170: 4296 – 4300, 2003.
42. Chambers RC, Leoni P, Blanc-Brude OP, Wembridge DE, and
Laurent GJ. Thrombin is a potent inducer of connective tissue
growth factor production via proteolytic activation of proteaseactivated receptor-1. J Biol Chem 275: 35584 –35591, 2000.
43. Chen J, Ishii M, Wang L, Ishii K, and Coughlin SR. 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.
44. Chen X, Earley K, Luo W, Lin SH, and Schilling WP. Functional
expression of a human thrombin receptor in Sf9 insect cells: evidence for an active tethered ligand. Biochem J 314: 603– 611, 1996.
45. Chen YH, Pouyssegur J, Courtneidge SA, and Van ObberghenSchilling E. Activation of Src family kinase activity by the G
protein-coupled thrombin receptor in growth-responsive fibroblasts. J Biol Chem 269: 27372–27377, 1994.
46. Cheng T, Liu D, Griffin JH, Fernandez JA, Castellino F,
Rosen ED, Fukudome K, and Zlokovic BV. Activated protein C
blocks p53-mediated apoptosis in ischemic human brain endothelium and is neuroprotective. Nat Med 9: 338 –342, 2003.
47. Cheung WM, Andrade-Gordon P, Derian CK, and Damiano BP.
Receptor-activating peptides distinguish thrombin receptor (PAR-1)
and protease activated receptor 2 (PAR-2) mediated hemodynamic
responses in vivo. Can J Physiol Pharmacol 76: 16 –25, 1998.
48. Cheung WM, D’Andrea MR, Andrade-Gordon P, and Damiano
BP. Altered vascular injury responses in mice deficient in proteaseactivated receptor-1. Arterioscler Thromb Vasc Biol 19: 3014 –3024,
1999.
49. Chow JM, Moffatt JD, and Cocks TM. Effect of protease-activated
receptor (PAR)-1, -2 and -4-activating peptides, thrombin and trypsin
in rat isolated airways. Br J Pharmacol 131: 1584–1591, 2000.
50. Cicala C. Protease activated receptor 2 and the cardiovascular
system. Br J Pharmacol 135: 14 –20, 2002.
51. Cicala C, Bucci M, De Dominicis G, Harriot P, Sorrentino L,
and Cirino G. Bronchoconstrictor effect of thrombin and throm-
PROTEASE-ACTIVATED RECEPTORS
73.
74.
75.
76.
78.
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
89.
Physiol Rev • VOL
90.
91.
92.
93.
94.
95.
96.
97.
98.
99.
100.
101.
102.
103.
104.
105.
106.
107.
108.
109.
ase-activated receptors: novel mechanisms of signaling by serine
proteases. Am J Physiol Cell Physiol 274: C1429 –C1452, 1998.
Dery O, Thoma MS, Wong H, Grady EF, and Bunnett NW.
Trafficking of proteinase-activated receptor-2 and beta-arrestin-1
tagged with green fluorescent protein. beta-Arrestin-dependent endocytosis of a proteinase receptor. J Biol Chem 274: 18524 –18535,
1999.
Dihanich M, Kaser M, Reinhard E, Cunningham D, and Monard D. Prothrombin mRNA is expressed by cells of the nervous
system. Neuron 6: 575–581, 1991.
Donovan FM and Cunningham DD. Signaling pathways involved
in thrombin-induced cell protection. J Biol Chem 273: 12746 –
12752, 1998.
Donovan FM, Pike CJ, Cotman CW, and Cunningham DD.
Thrombin induces apoptosis in cultured neurons and astrocytes via
a pathway requiring tyrosine kinase and RhoA activities. J Neurosci 17: 5316 –5326, 1997.
Ducroc R, Bontemps C, Marazova K, Devaud H, Darmoul D,
and Laburthe M. Trypsin is produced by and activates proteaseactivated receptor-2 in human cancer colon cells: evidence for new
autocrine loop. Life Sci 70: 1359 –1367, 2002.
Dulon S, Cande C, Bunnett NW, Hollenberg MD, Chignard M,
and Pidard D. Proteinase-activated receptor-2 and human lung
epithelial cells: disarming by neutrophil serine proteinases. Am J
Respir Cell Mol Biol 28: 339 –346, 2003.
Ekholm IE, Brattsand M, and Egelrud T. Stratum corneum
tryptic enzyme in normal epidermis: a missing link in the desquamation process? J Invest Dermatol 114: 56 – 63, 2000.
Even-Ram S, Uziely B, Cohen P, Grisaru-Granovsky S, Maoz
M, Ginzburg Y, Reich R, Vlodavsky I, and Bar-Shavit R. Thrombin receptor overexpression in malignant and physiological invasion processes. Nat Med 4: 909 –914, 1998.
Faruqi TR, Weiss EJ, Shapiro MJ, Huang W, and Coughlin SR.
Structure-function analysis of protease-activated receptor 4 tethered ligand peptides. Determinants of specificity and utility in
assays of receptor function. J Biol Chem 275: 19728 –19734, 2000.
Fee JA, Monsey JD, Handler RJ, Leonis MA, Mullaney SR,
Hope HM, and Silbert DF. 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.
Feng DM, Veber DF, Connolly TM, Condra C, Tang MJ, and
Nutt RF. Development of a potent thrombin receptor ligand. J Med
Chem 38: 4125– 4130, 1995.
Ferrell WR, Lockhart JC, Kelso EB, Dunning L, Plevin R,
Meek SE, Smith AJ, Hunter GD, McLean JS, McGarry F,
Ramage R, Jiang L, Kanke T, and Kawagoe J. Essential role for
proteinase-activated receptor-2 in arthritis. J Clin Invest 111: 35–
41, 2003.
Festoff BW, Smirnova IV, Ma J, and Citron BA. Thrombin, its
receptor and protease nexin I, its potent serpin, in the nervous
system. Semin Thromb Hemost 22: 267–271, 1996.
Fiorucci S, Mencarelli A, Palazzetti B, Distrutti E, Vergnolle
N, Hollenberg MD, Wallace JL, Morelli A, and Cirino G. Proteinase-activated receptor 2 is an anti-inflammatory signal for colonic lamina propria lymphocytes in a mouse model of colitis. Proc
Natl Acad Sci USA 98: 13936 –13941, 2001.
Gao C, Liu S, Hu HZ, Gao N, Kim GY, Xia Y, and Wood JD.
Serine proteases excite myenteric neurons through protease-activated receptors in guinea pig small intestine. Gastroenterology 123:
1554 –1564, 2002.
Gerszten RE, Chen J, Ishii M, Ishii K, Wang L, Nanevicz T,
Turck CW, Vu TK, and Coughlin SR. Specificity of the thrombin
receptor for agonist peptide is defined by its extracellular surface.
Nature 368: 648 – 651, 1994.
Gill JS, Pitts K, Rusnak FM, Owen WG, and Windebank AJ.
Thrombin induced inhibition of neurite outgrowth from dorsal root
ganglion neurons. Brain Res 797: 321–327, 1998.
Gingrich MB, Junge CE, Lyuboslavsky P, and Traynelis SF.
Potentiation of NMDA receptor function by the serine protease
thrombin. J Neurosci 20: 4582– 4595, 2000.
Grand RJ, Turnell AS, and Grabham PW. Cellular consequences
of thrombin-receptor activation. Biochem J 313: 353–368, 1996.
Green BT, Bunnett NW, Kulkarni-Narla A, Steinhoff M, and
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
77.
cin-based intervention of thrombin-receptor signaling and systemic
platelet activation. Nat Med 8: 1161–1165, 2002.
Cuffe JE, Bertog M, Velazquez-Rocha S, Dery O, Bunnett N,
and Korbmacher C. Basolateral PAR-2 receptors mediate KCl
secretion and inhibition of Na⫹ absorption in the mouse distal
colon. J Physiol 539: 209 –222, 2002.
Cumashi A, Ansuini H, Celli N, De Blasi A, O’Brien PJ, Brass
LF, and Molino M. Neutrophil proteases can inactivate human
PAR3 and abolish the co-receptor function of PAR3 on murine
platelets. Thromb Haemostasis 85: 533–538, 2001.
Cunningham MA, Rondeau E, Chen X, Coughlin SR, Holdsworth SR, and Tipping PG. Protease-activated receptor 1 mediates thrombin-dependent, cell-mediated renal inflammation in crescentic glomerulonephritis. J Exp Med 191: 455– 462, 2000.
Daaka Y, Luttrell LM, Ahn S, Della Rocca GJ, Ferguson SS,
Caron MG, and Lefkowitz RJ. Essential role for G proteincoupled receptor endocytosis in the activation of mitogen-activated protein kinase. J Biol Chem 273: 685– 688, 1998.
Damiano BP, Cheung WM, Santulli RJ, Fung-Leung WP, Ngo
K, Ye RD, Darrow AL, Derian CK, de Garavilla L, and Andrade-Gordon P. Cardiovascular responses mediated by proteaseactivated receptor-2 (PAR- 2) and thrombin receptor (PAR-1) are
distinguished in mice deficient in PAR-2 or PAR-1. J Pharmacol
Exp Ther 288: 671– 678, 1999.
Danahay H, Withey L, Poll CT, van de Graaf SF, and Bridges
RJ. Protease-activated receptor-2-mediated inhibition of ion transport in human bronchial epithelial cells. Am J Physiol Cell Physiol
280: C1455–C1464, 2001.
D’Andrea MR, Derian CK, Leturcq D, Baker SM, Brunmark A,
Ling P, Darrow AL, Santulli RJ, Brass LF, and AndradeGordon P. Characterization of protease-activated receptor-2 immunoreactivity in normal human tissues. J Histochem Cytochem
46: 157–164, 1998.
D’Andrea MR, Derian CK, Santulli RJ, and Andrade-Gordon
P. Differential expression of protease-activated receptors-1 and -2
in stromal fibroblasts of normal, benign, and malignant human
tissues. Am J Pathol 158: 2031–2041, 2001.
Darmoul D, Gratio V, Devaud H, Lehy T, and Laburthe M.
Aberrant expression and activation of the thrombin receptor protease-activated receptor-1 induces cell proliferation and motility in
human colon cancer cells. Am J Pathol 162: 1503–1513, 2003.
Darmoul D, Marie JC, Devaud H, Gratio V, and Laburthe M.
Initiation of human colon cancer cell proliferation by trypsin acting
at protease-activated receptor-2. Br J Cancer 85: 772–779, 2001.
Darrow AL, Fung-Leung WP, Ye RD, Santulli RJ, Cheung WM,
Derian CK, Burns CL, Damiano BP, Zhou L, Keenan CM,
Peterson PA, and Andrade-Gordon P. Biological consequences
of thrombin receptor deficiency in mice. Thromb Haemostasis 76:
860 – 866, 1996.
De Candia E, Hall SW, Rutella S, Landolfi R, Andrews RK, and
De Cristofaro R. Binding of thrombin to glycoprotein Ib accelerates the hydrolysis of Par-1 on intact platelets. J Biol Chem 276:
4692– 4698, 2001.
Defea K, Schmidlin F, Dery O, Grady EF, and Bunnett NW.
Mechanisms of initiation and termination of signalling by neuropeptide receptors: a comparison with the proteinase-activated
receptors. Biochem Soc Trans 28: 419 – 426, 2000.
DeFea KA, Zalevsky J, Thoma MS, Dery O, Mullins RD, and
Bunnett NW. Beta-arrestin-dependent endocytosis of proteinaseactivated receptor 2 is required for intracellular targeting of activated ERK1/2. J Cell Biol 148: 1267–1281, 2000.
De Garavilla L, Vergnolle N, Young SH, Ennes H, Steinhoff M,
Ossovskaya VS, D’Andrea MR, Mayer EA, Wallace JL, Hollenberg MD, Andrade-Gordon P, and Bunnett NW. Agonists of
proteinase-activated receptor 1 induce plasma extravasation by a
neurogenic mechanism. Br J Pharmacol 133: 975–987, 2001.
Derian CK, Damiano BP, Addo MF, Darrow AL, D’Andrea MR,
Nedelman M, Zhang HC, Maryanoff BE, and Andrade-Gordon
P. Blockade of the thrombin receptor protease-activated receptor-1
with a small-molecule antagonist prevents thrombus formation and
vascular occlusion in nonhuman primates. J Pharmacol Exp Ther
304: 855– 861, 2003.
Dery O, Corvera CU, Steinhoff M, and Bunnett NW. Protein-
615
616
110.
111.
112.
113.
114.
116.
117.
118.
119.
120.
121.
122.
123.
124.
125.
126.
127.
Brown DR. Intestinal type 2 proteinase-activated receptors: expression in opioid-sensitive secretomotor neural circuits that mediate epithelial ion transport. J Pharmacol Exp Ther 295: 410 – 416,
2000.
Griffin CT, Srinivasan Y, Zheng YW, Huang W, and Coughlin
SR. A role for thrombin receptor signaling in endothelial cells
during embryonic development. Science 293: 1666 –1670, 2001.
Griffin JH, Zlokovic B, and Fernandez JA. Activated protein C:
potential therapy for severe sepsis, thrombosis, and stroke. Semin
Hematol 39: 197–205, 2002.
Gurwitz D and Cunningham DD. Thrombin modulates and reverses neuroblastoma neurite outgrowth. Proc Natl Acad Sci USA
85: 3440 –3444, 1988.
Hamilton JR, Chow JM, and Cocks TM. Protease-activated receptor-2 turnover stimulated independently of receptor activation
in porcine coronary endothelial cells. Br J Pharmacol 127: 617–
622, 1999.
Hamilton JR and Cocks TM. Heterogeneous mechanisms of endothelium-dependent relaxation for thrombin and peptide activators of protease-activated receptor-1 in porcine isolated coronary
artery. Br J Pharmacol 130: 181–188, 2000.
Hamilton JR, Frauman AG, and Cocks TM. Increased expression of protease-activated receptor-2 (par2) and par4 in human
coronary artery by inflammatory stimuli unveils endothelium- dependent relaxations to par2 and par4 agonists. Circ Res 89: 92–98,
2001.
Hamilton JR, Moffatt JD, Frauman AG, and Cocks TM. Protease-activated receptor (PAR) 1 but not PAR2 or PAR4 mediates
endothelium-dependent relaxation to thrombin and trypsin in human pulmonary arteries. J Cardiovasc Pharmacol 38: 108 –119,
2001.
Hamilton JR, Moffatt JD, Tatoulis J, and Cocks TM. Enzymatic activation of endothelial protease-activated receptors is dependent on artery diameter in human and porcine isolated coronary arteries. Br J Pharmacol 136: 492–501, 2002.
Hayes KL and Tracy PB. The platelet high affinity binding site for
thrombin mimics hirudin, modulates thrombin-induced platelet activation, and is distinct from the glycoprotein Ib-IX-V complex.
J Biol Chem 274: 972–980, 1999.
He S, Peng Q, and Walls AF. Potent induction of a neutrophil and
eosinophil-rich infiltrate in vivo by human mast cell tryptase: selective enhancement of eosinophil recruitment by histamine. J Immunol 159: 6216 – 6225, 1997.
He S and Walls AF. Human mast cell tryptase: a stimulus of
microvascular leakage and mast cell activation. Eur J Pharmacol
328: 89 –97, 1997.
Hein L, Ishii K, Coughlin SR, and Kobilka BK. 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.
Henrikson KP, Salazar SL, Fenton JW II, and Pentecost BT.
Role of thrombin receptor in breast cancer invasiveness. Br J
Cancer 79: 401– 406, 1999.
Hollenberg MD, Saifeddine M, al-Ani B, and Kawabata A.
Proteinase-activated receptors: structural requirements for activity, receptor cross-reactivity, and receptor selectivity of receptoractivating peptides. Can J Physiol Pharmacol 75: 832– 841, 1997.
Hoogerwerf WA, Zou L, Shenoy M, Sun D, Micci MA, LeeHellmich H, Xiao SY, Winston JH, and Pasricha PJ. The proteinase-activated receptor 2 is involved in nociception. J Neurosci
21: 9036 –9042, 2001.
Horvat R and Palade GE. 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.
Hou L, Kapas S, Cruchley AT, Macey MG, Harriott P, Chinni
C, Stone SR, and Howells GL. Immunolocalization of proteaseactivated receptor-2 in skin: receptor activation stimulates interleukin-8 secretion by keratinocytes in vitro. Immunology 94: 356 –
362, 1998.
Howell DC, Laurent GJ, and Chambers RC. Role of thrombin
and its major cellular receptor, protease-activated receptor-1, in
pulmonary fibrosis. Biochem Soc Trans 30: 211–216, 2002.
Physiol Rev • VOL
128. Howells GL, Macey MG, Chinni C, Hou L, Fox MT, Harriott P,
and Stone SR. Proteinase-activated receptor-2: expression by human neutrophils. J Cell Sci 110: 881– 887, 1997.
129. Hoxie JA, Ahuja M, Belmonte E, Pizarro S, Parton R, and
Brass LF. Internalization and recycling of activated thrombin receptors. J Biol Chem 268: 13756 –13763, 1993.
130. Huang C, De Sanctis GT, O’Brien PJ, Mizgerd JP, Friend DS,
Drazen JM, Brass LF, and Stevens RL. Evaluation of the substrate specificity of human mast cell tryptase beta I and demonstration of its importance in bacterial infections of the lung. J Biol
Chem 276: 26276 –26284, 2001.
131. Hung DT, Vu TH, Nelken NA, and Coughlin SR. Thrombininduced events in non-platelet cells are mediated by the unique
proteolytic mechanism established for the cloned platelet thrombin
receptor. J Cell Biol 116: 827– 832, 1992.
132. Hung DT, Wong YH, Vu TK, and Coughlin SR. 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.
133. Iaccarino G, Rockman HA, Shotwell KF, Tomhave ED, and
Koch WJ. Myocardial overexpression of GRK3 in transgenic mice:
evidence for in vivo selectivity of GRKs. Am J Physiol Heart Circ
Physiol 275: H1298 –H1306, 1998.
134. Idell S, Gonzalez K, Bradford H, MacArthur CK, Fein AM,
Maunder RJ, Garcia JG, Griffith DE, Weiland J, and Martin
TR. Procoagulant activity in bronchoalveolar lavage in the adult
respiratory distress syndrome. Contribution of tissue factor associated with factor VII. Am Rev Respir Dis 136: 1466 –1474, 1987.
135. Ishihara H, Connolly AJ, Zeng D, Kahn ML, Zheng YW, Timmons C, Tram T, and Coughlin SR. Protease-activated receptor
3 is a second thrombin receptor in humans. Nature 386: 502–506,
1997.
136. Ishii K, Chen J, Ishii M, Koch WJ, Freedman NJ, Lefkowitz
RJ, and Coughlin SR. 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.
137. Ishii K, Hein L, Kobilka B, and Coughlin SR. Kinetics of thrombin receptor cleavage on intact cells. Relation to signaling. J Biol
Chem 268: 9780 –9786, 1993.
138. Jalink K, van Corven EJ, Hengeveld T, Morii N, Narumiya S,
and Moolenaar WH. 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.
139. Jarjour NN, Calhoun WJ, Schwartz LB, and Busse WW. Elevated bronchoalveolar lavage fluid histamine levels in allergic asthmatics are associated with increased airway obstruction. Am Rev
Respir Dis 144: 83– 87, 1991.
140. Jin E, Fujiwara M, Pan X, Ghazizadeh M, Arai S, Ohaki Y,
Kajiwara K, Takemura T, and Kawanami O. Protease-activated
receptor (PAR)-1 and PAR-2 participate in the cell growth of alveolar capillary endothelium in primary lung adenocarcinomas. Cancer 97: 703–713, 2003.
141. Julius D and Basbaum AI. Molecular mechanisms of nociception.
Nature 413: 203–210, 2001.
142. Kahn ML, Nakanishi-Matsui M, Shapiro MJ, Ishihara H, and
Coughlin SR. Protease-activated receptors 1 and 4 mediate activation of human platelets by thrombin. J Clin Invest 103: 879 – 887,
1999.
143. Kahn ML, Zheng YW, Huang W, Bigornia V, Zeng D, Moff S,
Farese RV Jr, Tam C, and Coughlin SR. A dual thrombin receptor system for platelet activation. Nature 394: 690 – 694, 1998.
144. Kamath L, Meydani A, Foss F, and Kuliopulos A. Signaling
from protease-activated receptor-1 inhibits migration and invasion
of breast cancer cells. Cancer Res 61: 5933–5940, 2001.
145. Kanke T, Macfarlane SR, Seatter MJ, Davenport E, Paul A,
McKenzie R, and Plevin R. Proteinase-activated receptor-2-mediated activation of stress-activated protein kinases and inhibitory
kappa B kinases in NCTC 2544 keratinocytes. J Biol Chem 18: 18,
2001.
146. Katona G, Berglund GI, Hajdu J, Graf L, and Szilagyi L.
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
115.
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
PROTEASE-ACTIVATED RECEPTORS
147.
148.
149.
150.
151.
153.
154.
155.
156.
157.
158.
159.
160.
161.
162.
163.
164.
Physiol Rev • VOL
165.
166.
167.
168.
169.
170.
171.
172.
173.
174.
175.
176.
177.
178.
179.
180.
181.
182.
183.
airways: upregulation of PAR-2 in respiratory epithelium from patients with asthma. J Allergy Clin Immunol 108: 797– 803, 2001.
Koivunen E, Huhtala ML, and Stenman UH. Human ovarian
tumor-associated trypsin. Its purification and characterization from
mucinous cyst fluid and identification as an activator of pro-urokinase. J Biol Chem 264: 14095–14099, 1989.
Koivunen E, Saksela O, Itkonen O, Osman S, Huhtala ML, and
Stenman UH. Human colonic carcinoma, fibrosarcoma and leukemia cell lines produce tumor-associated trypsinogen. Int J Cancer
47: 592–596, 1991.
Kong W, McConalogue K, Khitin LM, Hollenberg MD, Payan
DG, Bohm SK, and Bunnett NW. Luminal trypsin may regulate
enterocytes through proteinase-activated receptor 2. Proc Natl
Acad Sci USA 94: 8884 – 8889, 1997.
Koshikawa N, Hasegawa S, Nagashima Y, Mitsuhashi K, Tsubota Y, Miyata S, Miyagi Y, Yasumitsu H, and Miyazaki K.
Expression of trypsin by epithelial cells of various tissues, leukocytes, and neurons in human and mouse. Am J Pathol 153: 937–944,
1998.
Koshikawa N, Nagashima Y, Miyagi Y, Mizushima H, Yanoma
S, Yasumitsu H, and Miyazaki K. Expression of trypsin in vascular endothelial cells. FEBS Lett 409: 442– 448, 1997.
Krishna MT, Chauhan A, Little L, Sampson K, Hawksworth R,
Mant T, Djukanovic R, Lee T, and Holgate S. Inhibition of mast
cell tryptase by inhaled APC 366 attenuates allergen-induced latephase airway obstruction in asthma. J Allergy Clin Immunol 107:
1039 –1045, 2001.
Kunzelmann K, Schreiber R, Konig J, and Mall M. Ion transport
induced by proteinase-activated receptors (PAR2) in colon and
airways. Cell Biochem Biophys 36: 209 –214, 2002.
Lan RS, Knight DA, Stewart GA, and Henry PJ. Role of PGE(2)
in protease-activated receptor-1, -2 and -4 mediated relaxation in
the mouse isolated trachea. Br J Pharmacol 132: 93–100, 2001.
Lan RS, Stewart GA, and Henry PJ. Modulation of airway
smooth muscle tone by protease activated receptor- 1,-2,-3 and -4 in
trachea isolated from influenza A virus-infected mice. Br J Pharmacol 129: 63–70, 2000.
Laposata M, Dovnarsky DK, and Shin HS. Thrombin-induced
gap formation in confluent endothelial cell monolayers in vitro.
Blood 62: 549 –556, 1983.
Lerner DJ, Chen M, Tram T, and Coughlin SR. 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.
Linden DR, Manning BP, Bunnett NW, and Mawe GM. Agonists
of proteinase-activated receptor 2 excite guinea pig ileal myenteric
neurons. Eur J Pharmacol 431: 311–314, 2001.
Lindner JR, Kahn ML, Coughlin SR, Sambrano GR, Schauble
E, Bernstein D, Foy D, Hafezi-Moghadam A, and Ley K. Delayed onset of inflammation in protease-activated receptor-2-deficient mice. J Immunol 165: 6504 – 6510, 2000.
Lourbakos A, Chinni C, Thompson P, Potempa J, Travis J,
Mackie EJ, and Pike RN. Cleavage and activation of proteinaseactivated receptor-2 on human neutrophils by gingipain-R from
Porphyromonas gingivalis. FEBS Lett 435: 45– 48, 1998.
Lourbakos A, Potempa J, Travis J, D’Andrea MR, AndradeGordon P, Santulli R, Mackie EJ, and Pike RN. Arginine-specific protease from Porphyromonas gingivalis activates proteaseactivated receptors on human oral epithelial cells and induces
interleukin-6 secretion. Infect Immun 69: 5121–5130., 2001.
Lourbakos A, Yuan YP, Jenkins AL, Travis J, Andrade-Gordon P, Santulli R, Potempa J, and Pike RN. Activation of
protease-activated receptors by gingipains from Porphyromonas
gingivalis leads to platelet aggregation: a new trait in microbial
pathogenicity. Blood 97: 3790 –3797, 2001.
Luttrell LM and Lefkowitz RJ. The role of beta-arrestins in the
termination and transduction of G-protein-coupled receptor signals. J Cell Sci 115: 455– 465, 2002.
Macfarlane SR, Seatter MJ, Kanke T, Hunter GD, and Plevin
R. Proteinase-activated receptors. Pharmacol Rev 53: 245–282,
2001.
Madamanchi NR, Li S, Patterson C, and Runge MS. Thrombin
regulates vascular smooth muscle cell growth and heat shock
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
152.
Crystal structure reveals basis for the inhibitor resistance of human
brain trypsin. J Mol Biol 315: 1209 –1218, 2002.
Kauffman HF, Tomee JF, van de Riet MA, Timmerman AJ,
and Borger P. Protease-dependent activation of epithelial cells by
fungal allergens leads to morphologic changes and cytokine production. J Allergy Clin Immunol 105: 1185–1193, 2000.
Kaufmann R, Patt S, Zieger M, Kraft R, Tausch S, Henklein P,
and Nowak G. The two-receptor system PAR-1/PAR-4 mediates
alpha-thrombin-induced [Ca2⫹]i mobilization in human astrocytoma cells. J Cancer Res Clin Oncol 126: 91–94, 2000.
Kawabata A, Kawao N, Kuroda R, Itoh H, and Nishikawa H.
Specific expression of spinal Fos after PAR-2 stimulation in mast
cell-depleted rats. Neuroreport 13: 511–514, 2002.
Kawabata A, Kawao N, Kuroda R, Tanaka A, Itoh H, and
Nishikawa H. Peripheral PAR-2 triggers thermal hyperalgesia and
nociceptive responses in rats. Neuroreport 12: 715–719, 2001.
Kawabata A, Kawao N, Kuroda R, Tanaka A, and Shimada C.
The PAR-1-activating peptide attenuates carrageenan-induced hyperalgesia in rats. Peptides 23: 1181–1183, 2002.
Kawabata A, Kinoshita M, Kuroda R, and Kakehi K. Capsazepine partially inhibits neurally mediated gastric mucus secretion
following activation of protease-activated receptor 2. Clin Exp
Pharmacol Physiol 29: 360 –361, 2002.
Kawabata A, Kinoshita M, Nishikawa H, Kuroda R, Nishida M,
Araki H, Arizono N, Oda Y, and Kakehi K. The protease-activated receptor-2 agonist induces gastric mucus secretion and mucosal cytoprotection. J Clin Invest 107: 1443–1450, 2001.
Kawabata A, Kuroda R, Kuroki N, Nishikawa H, and Kawai K.
Dual modulation by thrombin of the motility of rat oesophageal
muscularis mucosae via two distinct protease-activated receptors
(PARs): a novel role for PAR-4 as opposed to PAR-1. Br J Pharmacol 131: 578 –584, 2000.
Kawabata A, Kuroda R, Nagata N, Kawao N, Masuko T, Nishikawa H, and Kawai K. In vivo evidence that protease-activated
receptors 1 and 2 modulate gastrointestinal transit in the mouse.
Br J Pharmacol 133: 1213–1218, 2001.
Kawabata A, Kuroda R, Nishida M, Nagata N, Sakaguchi Y,
Kawao N, Nishikawa H, Arizono N, and Kawai K. Proteaseactivated receptor-2 (PAR-2) in the pancreas and parotid gland:
immunolocalization and involvement of nitric oxide in the evoked
amylase secretion. Life Sci 71: 2435–2446, 2002.
Kawabata A, Morimoto N, Nishikawa H, Kuroda R, Oda Y, and
Kakehi K. Activation of protease-activated receptor-2 (PAR-2)
triggers mucin secretion in the rat sublingual gland. Biochem Biophys Res Commun 270: 298 –302, 2000.
Kawabata A, Nishikawa H, Kuroda R, Kawai K, and Hollenberg MD. Proteinase-activated receptor-2 (PAR-2): regulation of
salivary and pancreatic exocrine secretion in vivo in rats and mice.
Br J Pharmacol 129: 1808 –1814, 2000.
Kawagoe J, Takizawa T, Matsumoto J, Tamiya M, Meek SE,
Smith AJ, Hunter GD, Plevin R, Saito N, Kanke T, Fujii M, and
Wada Y. Effect of protease-activated receptor-2 deficiency on allergic dermatitis in the mouse ear. Jpn J Pharmacol 88: 77– 84,
2002.
Kawao N, Sakaguchi Y, Tagome A, Kuroda R, Nishida S, Irimajiri K, Nishikawa H, Kawai K, Hollenberg MD, and Kawabata A. Protease-activated receptor-2 (PAR-2) in the rat gastric
mucosa: immunolocalization and facilitation of pepsin/pepsinogen
secretion. Br J Pharmacol 135: 1292–1296, 2002.
Keogh RJ, Houliston RA, and Wheeler-Jones CP. Thrombinstimulated Pyk2 phosphorylation in human endothelium is dependent on intracellular calcium and independent of protein kinase C
and Src kinases. Biochem Biophys Res Commun 294: 1001–1008,
2002.
King C, Brennan S, Thompson PJ, and Stewart GA. Dust mite
proteolytic allergens induce cytokine release from cultured airway
epithelium. J Immunol 161: 3645–3651, 1998.
Klages B, Brandt U, Simon MI, Schultz G, and Offermanns S.
Activation of G12/G13 results in shape change and Rho/Rho-kinasemediated myosin light chain phosphorylation in mouse platelets.
J Cell Biol 144: 745–754, 1999.
Knight DA, Lim S, Scaffidi AK, Roche N, Chung KF, Stewart
GA, and Thompson PJ. Protease-activated receptors in human
617
618
184.
185.
186.
187.
188.
190.
191.
192.
193.
194.
195.
196.
197.
198.
199.
200.
201.
202.
proteins via the JAK-STAT pathway. J Biol Chem 276: 18915–18924,
2001.
Magazine HI, King JM, and Srivastava KD. Protease activated
receptors modulate aortic vascular tone. Int J Cardiol 53 Suppl:
S75–S80, 1996.
Marchese A and Benovic JL. Agonist-promoted ubiquitination of
the G protein-coupled receptor CXCR4 mediates lysosomal sorting.
J Biol Chem 276: 45509 – 45512, 2001.
Marty I, Peclat V, Kirdaite G, Salvi R, So A, and Busso N.
Amelioration of collagen-induced arthritis by thrombin inhibition.
J Clin Invest 107: 631– 640, 2001.
Maryanoff BE, Santulli RJ, McComsey DF, Hoekstra WJ,
Hoey K, Smith CE, Addo M, Darrow AL, and Andrade-Gordon
P. Protease-activated receptor-2 (PAR-2): structure-function study
of receptor activation by diverse peptides related to tethered-ligand
epitopes. Arch Biochem Biophys 386: 195–204, 2001.
McEuen AR, He S, Brander ML, and Walls AF. Guinea pig lung
tryptase. Localisation to mast cells and characterisation of the
partially purified enzyme. Biochem Pharmacol 52: 331–340, 1996.
McNamara CA, Sarembock IJ, Gimple LW, Fenton JW,
Coughlin SR, and Owens GK. Thrombin stimulates proliferation
of cultured rat aortic smooth muscle cells by a proteolytically
activated receptor. J Clin Invest 91: 94 –98, 1993.
Metcalfe DD, Baram D, and Mekori YA. Mast cells. Physiol Rev
77: 1033–1079, 1997.
Milia AF, Salis MB, Stacca T, Pinna A, Madeddu P, Trevisani
M, Geppetti P, and Emanueli C. Protease-activated receptor-2
stimulates angiogenesis and accelerates hemodynamic recovery in
a mouse model of hindlimb ischemia. Circ Res 91: 346 –352, 2002.
Miotto D, Hollenberg MD, Bunnett NW, Papi A, Braccioni F,
Boschetto P, Rea F, Zuin A, Geppetti P, Saetta M, Maestrelli
P, Fabbri LM, and Mapp CE. Expression of protease activated
receptor-2 (PAR-2) in central airways of smokers and non-smokers.
Thorax 57: 146 –151, 2002.
Mirza H, Schmidt VA, Derian CK, Jesty J, and Bahou WF.
Mitogenic responses mediated through the proteinase-activated
receptor-2 are induced by expressed forms of mast cell alpha- or
beta-tryptases. Blood 90: 3914 –3922, 1997.
Mirza H, Yatsula V, and Bahou WF. The proteinase activated
receptor-2 (PAR-2) mediates mitogenic responses in human vascular endothelial cells. J Clin Invest 97: 1705–1714, 1996.
Moffatt JD, Jeffrey KL, and Cocks TM. Protease-activated receptor-2 activating peptide SLIGRL inhibits bacterial lipopolysaccharide-induced recruitment of polymorphonuclear leukocytes
into the airways of mice. Am J Respir Cell Mol Biol 26: 680 – 684,
2002.
Molino M, Barnathan ES, Numerof R, Clark J, Dreyer M,
Cumashi A, Hoxie JA, Schechter N, Woolkalis M, and Brass
LF. Interactions of mast cell tryptase with thrombin receptors and
PAR-2. J Biol Chem 272: 4043– 4049, 1997.
Molino M, Blanchard N, Belmonte E, Tarver AP, Abrams C,
Hoxie JA, Cerletti C, and Brass LF. Proteolysis of the human
platelet and endothelial cell thrombin receptor by neutrophil-derived cathepsin G. J Biol Chem 270: 11168 –11175, 1995.
Mule F, Baffi MC, and Cerra MC. Dual effect mediated by
protease-activated receptors on the mechanical activity of rat colon. Br J Pharmacol 136: 367–374, 2002.
Nakanishi-Matsui M, Zheng YW, Sulciner DJ, Weiss EJ, Ludeman MJ, and Coughlin SR. PAR3 is a cofactor for PAR4
activation by thrombin. Nature 404: 609 – 613, 2000.
Nakayama T, Hirano K, Shintani Y, Nishimura J, Nakatsuka
A, Kuga H, Takahashi S, and Kanaide H. Unproductive cleavage
and the inactivation of protease-activated receptor-1 by trypsin in
vascular endothelial cells. Br J Pharmacol 138: 121–130, 2003.
Nanevicz T, Ishii M, Wang L, Chen M, Chen J, Turck CW,
Cohen FE, and Coughlin SR. Mechanisms of thrombin receptor
agonist specificity. Chimeric receptors and complementary mutations identify an agonist recognition site. J Biol Chem 270: 21619 –
21625, 1995.
Nanevicz T, Wang L, Chen M, Ishii M, and Coughlin SR.
Thrombin receptor activating mutations. Alteration of an extracellular agonist recognition domain causes constitutive signaling.
J Biol Chem 271: 702–706, 1996.
Physiol Rev • VOL
203. Napoli C, Cicala C, Wallace JL, de Nigris F, Santagada V,
Caliendo G, Franconi F, Ignarro LJ, and Cirino G. Proteaseactivated receptor-2 modulates myocardial ischemia-reperfusion
injury in the rat heart. Proc Natl Acad Sci USA 97: 3678 –3683, 2000.
204. Napoli C, De Nigris F, Cicala C, Wallace JL, Caliendo G,
Condorelli M, Santagada V, and Cirino G. Protease-activated
receptor-2 activation improves efficiency of experimental ischemic
preconditioning. Am J Physiol Heart Circ Physiol 282: H2004 –
H2010, 2002.
205. Nguyen TD, Moody MW, Steinhoff M, Okolo C, Koh DS, and
Bunnett NW. Trypsin activates pancreatic duct epithelial cell ion
channels through proteinase-activated receptor-2. J Clin Invest
103: 261–269, 1999.
206. Niclou S, Suidan HS, Brown-Luedi M, and Monard D. Expression of the thrombin receptor mRNA in rat brain. Cell Mol Biol 40:
421– 428, 1994.
207. Niclou SP, Suidan HS, Pavlik A, Vejsada R, and Monard D.
Changes in the expression of protease-activated receptor 1 and
protease nexin-1 mRNA during rat nervous system development
and after nerve lesion. Eur J Neurosci 10: 1590 –1607, 1998.
208. Nierodzik ML, Bain RM, Liu LX, Shivji M, Takeshita K, and
Karpatkin S. 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.
209. Nierodzik ML, Chen K, Takeshita K, Li JJ, Huang YQ, Feng
XS, D’Andrea MR, Andrade-Gordon P, and Karpatkin S. Protease-activated receptor 1 (PAR-1) is required and rate-limiting for
thrombin-enhanced experimental pulmonary metastasis. Blood 92:
3694 –3700, 1998.
210. Nierodzik ML, Kajumo F, and Karpatkin S. 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.
211. Nishikawa H, Kawai K, Nishimura S, Tanaka S, Araki H,
Al-Ani B, Hollenberg MD, Kuroda R, and Kawabata A. Suppression by protease-activated receptor-2 activation of gastric acid
secretion in rats. Eur J Pharmacol 447: 87–90, 2002.
212. Norton KJ, Scarborough RM, Kutok JL, Escobedo MA, Nannizzi L, and Coller BS. 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.
213. Nystedt S, Emilsson K, Larsson AK, Strombeck B, and Sundelin J. Molecular cloning and functional expression of the gene
encoding the human proteinase-activated receptor 2. Eur J Biochem 232: 84 – 89, 1995.
214. Nystedt S, Emilsson K, Wahlestedt C, and Sundelin J. Molecular cloning of a potential proteinase activated receptor. Proc Natl
Acad Sci USA 91: 9208 –9212, 1994.
215. Nystedt S, Larsson AK, Aberg H, and Sundelin J. The mouse
proteinase-activated receptor-2 cDNA and gene. Molecular cloning
and functional expression. J Biol Chem 270: 5950 –5955, 1995.
216. Nystedt S, Ramakrishnan V, and Sundelin J. The proteinaseactivated receptor 2 is induced by inflammatory mediators in human endothelial cells. Comparison with the thrombin receptor.
J Biol Chem 271: 14910 –14915, 1996.
217. O’Brien PJ, Molino M, Kahn M, and Brass LF. Protease activated receptors: theme and variations. Oncogene 20: 1570 –1581,
2001.
218. O’Brien PJ, Prevost N, Molino M, Hollinger MK, Woolkalis
MJ, Woulfe DS, and Brass LF. Thrombin responses in human
endothelial cells. Contributions from receptors other than PAR1
include the transactivation of PAR2 by thrombin-cleaved PAR1.
J Biol Chem 275: 13502–13509, 2000.
219. Offermanns S. In vivo functions of heterotrimeric G-proteins:
studies in Galpha-deficient mice. Oncogene 20: 1635–1642, 2001.
220. Offermanns S, Laugwitz KL, Spicher K, and Schultz G. 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.
221. Offermanns S, Mancino V, Revel JP, and Simon MI. Vascular
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
189.
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
PROTEASE-ACTIVATED RECEPTORS
222.
223.
224.
225.
226.
228.
229.
230.
231.
232.
233.
234.
235.
236.
237.
238.
239.
Physiol Rev • VOL
240.
241.
242.
243.
244.
245.
246.
247.
248.
249.
250.
251.
252.
253.
254.
255.
256.
and initiation of coagulation by tissue factor. Proc Natl Acad Sci
USA 98: 7742–7747, 2001.
Riewald M and Ruf W. Science review: role of coagulation protease cascades in sepsis. Crit Care 7: 123–129, 2003.
Robin J, Kharbanda R, McLean P, Campbell R, and Vallance P.
Protease-activated receptor 2-mediated vasodilatation in humans
in vivo: role of nitric oxide and prostanoids. Circulation 107:
954 –959, 2003.
Roosterman D, Schmidlin F, and Bunnett NW. Rab5a and
rab11a mediate agonist-induced trafficking of protease-activated
receptor 2. Am J Physiol Cell Physiol 284: C1319 –C1329, 2003.
Roy SS, Saifeddine M, Loutzenhiser R, Triggle CR, and Hollenberg MD. Dual endothelium-dependent vascular activities of
proteinase-activated receptor-2-activating peptides: evidence for
receptor heterogeneity. Br J Pharmacol 123: 1434 –1440, 1998.
Ruoss S, Hartmann T, and Caughey G. Mast cell tryptase is a
mitogen for cultured fibroblasts. J Clin Invest 88: 493– 499, 1991.
Sabri A, Alcott SG, Elouardighi H, Pak E, Derian C, AndradeGordon P, Kinnally K, and Steinberg SF. Neutrophil cathepsin
G promotes detachment-induced cardiomyocyte apoptosis via a
protease-activated receptor-independent mechanism. J Biol Chem.
In press.
Sabri A, Short J, Guo J, and Steinberg SF. Protease-activated
receptor-1-mediated DNA synthesis in cardiac fibroblast is via epidermal growth factor receptor transactivation: distinct PAR-1 signaling pathways in cardiac fibroblasts and cardiomyocytes. Circ
Res 91: 532–539, 2002.
Saifeddine M, al-Ani B, Cheng CH, Wang L, and Hollenberg
MD. Rat proteinase-activated receptor-2 (PAR-2): cDNA sequence
and activity of receptor-derived peptides in gastric and vascular
tissue. Br J Pharmacol 118: 521–530, 1996.
Saifeddine M, Roy SS, Al-Ani B, Triggle CR, and Hollenberg
MD. Endothelium-dependent contractile actions of proteinase-activated receptor-2-activating peptides in human umbilical vein: release of a contracting factor via a novel receptor. Br J Pharmacol
125: 1445–1454, 1998.
Sambrano GR, Huang W, Faruqi T, Mahrus S, Craik C, and
Coughlin SR. Cathepsin G activates protease-activated receptor-4
in human platelets. J Biol Chem 275: 6819 – 6823, 2000.
Santos J, Bayarri C, Saperas E, Nogueiras C, Antolin M,
Mourelle M, Cadahia A, and Malagelada JR. Characterisation of
immune mediator release during the immediate response to segmental mucosal challenge in the jejunum of patients with food
allergy. Gut 45: 553–558, 1999.
Santulli RJ, Derian CK, Darrow AL, Tomko KA, Eckardt AJ,
Seiberg M, Scarborough RM, and Andrade-Gordon P. 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.
Sawada K, Nishibori M, Nakaya N, Wang Z, and Saeki K.
Purification and characterization of a trypsin-like serine proteinase
from rat brain slices that degrades laminin and type IV collagen and
stimulates protease-activated receptor-2. J Neurochem 74: 1731–
1738, 2000.
Scarborough RM, Naughton MA, Teng W, Hung DT, Rose J, Vu
TK, Wheaton VI, Turck CW, and Coughlin SR. 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.
Schechter NM, Brass LF, Lavker RM, and Jensen PJ. Reaction
of mast cell proteases tryptase and chymase with protease activated receptors (PARs) on keratinocytes and fibroblasts. J Cell
Physiol 176: 365–373, 1998.
Schmidlin F, Amadesi S, Dabbagh K, Lewis DE, Knott P,
Bunnett NW, Gater PR, Geppetti P, Bertrand C, and Stevens
ME. Protease-activated receptor 2 mediates eosinophil infiltration
and hyperreactivity in allergic inflammation of the airway. J Immunol 169: 5315–5321, 2002.
Schmidlin F, Amadesi S, Vidil R, Trevisani M, Martinet N,
Caughey G, Tognetto M, Cavallesco G, Mapp C, Geppetti P,
and Bunnett NW. Expression and function of proteinase-activated
receptor 2 in human bronchial smooth muscle. Am J Respir Crit
Care Med 164: 1276 –1281, 2001.
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
227.
system defects and impaired cell chemokinesis as a result of Galpha13 deficiency. Science 275: 533–536, 1997.
Offermanns S, Toombs CF, Hu YH, and Simon MI. Defective
platelet activation in G alpha(q)-deficient mice. Nature 389: 183–
186, 1997.
Ofosu FA, Freedman J, Dewar L, Song Y, and Fenton JW II. A
trypsin-like platelet protease propagates protease-activated receptor-1 cleavage and platelet activation. Biochem J 336: 283–285,
1998.
Paine C, Sharlow E, Liebel F, Eisinger M, Shapiro S, and
Seiberg M. An alternative approach to depigmentation by soybean
extracts via inhibition of the PAR-2 pathway. J Invest Dermatol
116: 587–595, 2001.
Paing MM, Stutts AB, Kohout TA, Lefkowitz RJ, and Trejo J.
beta-Arrestins regulate protease-activated receptor-1 desensitization but not internalization or down-regulation. J Biol Chem 277:
1292–1300, 2002.
Patterson C, Stouffer GA, Madamanchi N, and Runge MS.
New tricks for old dogs: nonthrombotic effects of thrombin in
vessel wall biology. Circ Res 88: 987–997, 2001.
Pendurthi UR, Ngyuen M, Andrade-Gordon P, Petersen LC,
and Rao LV. Plasmin induces Cyr61 gene expression in fibroblasts
via protease-activated receptor-1 and p44/42 mitogen-activated protein kinase-dependent signaling pathway. Arterioscler Thromb
Vasc Biol 22: 1421–1426, 2002.
Pereira PJ, Bergner A, Macedo-Ribeiro S, Huber R,
Matschiner G, Fritz H, Sommerhoff CP, and Bode W. Human
beta-tryptase is a ring-like tetramer with active sites facing a central pore. Nature 392: 306 –311, 1998.
Perraud F, Besnard F, Sensenbrenner M, and Labourdette G.
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.
Pindon A, Berry M, and Hantai D. Thrombomodulin as a new
marker of lesion-induced astrogliosis: involvement of thrombin
through the G-protein-coupled protease-activated receptor-1.
J Neurosci 20: 2543–2550, 2000.
Prenzel N, Zwick E, Daub H, Leserer M, Abraham R, Wallasch
C, and Ullrich A. EGF receptor transactivation by G-proteincoupled receptors requires metalloproteinase cleavage of proHBEGF. Nature 402: 884 – 888, 1999.
Raithel M, Winterkamp S, Pacurar A, Ulrich P, Hochberger J,
and Hahn EG. Release of mast cell tryptase from human colorectal mucosa in inflammatory bowel disease. Scand J Gastroenterol
36: 174 –179, 2001.
Rasmussen UB, Vouret-Craviari V, Jallat S, Schlesinger Y,
Pages G, Pavirani A, Lecocq JP, Pouyssegur J, and Van Obberghen-Schilling E. cDNA cloning and expression of a hamster
alpha-thrombin receptor coupled to Ca2⫹ mobilization. FEBS Lett
288: 123–128, 1991.
Rattenholl A and Steinhoff M. Role of proteinase-activated receptors in cutaneous biology and disease. In: Drug Development
Research, edited by M. Hollenberg and N. Vergrolle. New York:
Wiley-Liss, 2003, vol. 59, p. 408 – 417.
Reed DE, Barajas-Lopez C, Cottrell G, Velazquez-Rocha S,
Dery O, Grady EF, Bunnett NW, and Vanner SJ. Mast cell
tryptase and proteinase-activated receptor 2 induce hyperexcitability of guinea-pig submucosal neurons. J Physiol 547: 531–542, 2003.
Renesto P, Si-Tahar M, Moniatte M, Balloy V, Van Dorsselaer
A, Pidard D, and Chignard M. Specific inhibition of thrombininduced 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.
Ricciardolo FL, Steinhoff M, Amadesi S, Guerrini R, Tognetto
M, Trevisani M, Creminon C, Bertrand C, Bunnett NW, Fabbri
LM, Salvadori S, and Geppetti P. Presence and bronchomotor
activity of protease-activated receptor-2 in guinea pig airways.
Am J Respir Crit Care Med 161: 1672–1680, 2000.
Riewald M, Petrovan RJ, Donner A, Mueller BM, and Ruf W.
Activation of endothelial cell protease activated receptor 1 by the
protein C pathway. Science 296: 1880 –1882, 2002.
Riewald M and Ruf W. Mechanistic coupling of protease signaling
619
620
VALERIA S. OSSOVSKAYA AND NIGEL W. BUNNETT
Physiol Rev • VOL
275. Suzuki Y, Nomura J, Hori J, Koyama J, Takahashi M, and
Horii I. Detection and characterization of endogeneous protease
associated with desquamation of stratum corneum. Arch Dermatol
Res 285: 372–377, 1993.
276. Takeuchi T, Harris JL, Huang W, Yan KW, Coughlin SR, and
Craik CS. Cellular localization of membrane-type serine protease
1 and identification of protease-activated receptor-2 and singlechain urokinase-type plasminogen activator as substrates. J Biol
Chem 275: 26333–26342, 2000.
277. Tay-Uyboco J, Poon MC, Ahmad S, and Hollenberg MD. 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.
278. Tiruppathi C, Yan W, Sandoval R, Naqvi T, Pronin AN,
Benovic JL, and Malik AB. G protein-coupled receptor kinase-5
regulates thrombin-activated signaling in endothelial cells. Proc
Natl Acad Sci USA 97: 7440 –7445, 2000.
279. Toyoda N, Gabazza EC, Inoue H, Araki K, Nakashima S, Oka
S, Taguchi Y, Nakamura M, Suzuki Y, Taguchi O, Imoto I,
Suzuki K, and Adachi Y. Expression and cytoprotective effect of
protease-activated receptor-1 in gastric epithelial cells. Scand J
Gastroenterol 38: 253–259, 2003.
280. Tran T and Stewart AG. Protease-activated receptor (PAR)-independent growth and pro-inflammatory actions of thrombin on human cultured airway smooth muscle. Br J Pharmacol 138: 865– 875,
2003.
281. Trejo J, Altschuler Y, Fu HW, Mostov KE, and Coughlin SR.
Protease-activated receptor-1 down-regulation: a mutant HeLa cell
line suggests novel requirements for PAR1 phosphorylation and
recruitment to clathrin-coated pits. J Biol Chem 275: 31255–31265,
2000.
282. Trejo J and Coughlin SR. The cytoplasmic tails of proteaseactivated receptor-1 and substance P receptor specify sorting to
lysosomes versus recycling. J Biol Chem 274: 2216 –2224, 1999.
283. Trejo J, Hammes SR, and Coughlin SR. Termination of signaling
by protease-activated receptor-1 is linked to lysosomal sorting.
Proc Natl Acad Sci USA 95: 13698 –13702, 1998.
284. Tremaine WJ, Brzezinski A, Katz JA, Wolf DC, Fleming TJ,
Mordenti J, Strenkoski-Nix LC, and Kurth MC. Treatment of
mildly to moderately active ulcerative colitis with a tryptase inhibitor (APC 2059): an open-label pilot study. Aliment Pharmacol
Ther 16: 407– 413, 2002.
285. Tsopanoglou NE and Maragoudakis ME. On the mechanism of
thrombin-induced angiogenesis. Potentiation of vascular endothelial growth factor activity on endothelial cells by up-regulation of
its receptors. J Biol Chem 274: 23969 –23976, 1999.
286. Turgeon VL, Lloyd ED, Wang S, Festoff BW, and Houenou LJ.
Thrombin perturbs neurite outgrowth and induces apoptotic cell
death in enriched chick spinal motoneuron cultures through
caspase activation. J Neurosci 18: 6882– 6891, 1998.
287. Ubl JJ and Reiser G. Activity of protein kinase C is necessary for
sustained thrombin-induced [Ca2⫹]i oscillations in rat glioma cells.
Pflügers Arch 433: 312–320, 1997.
288. Ubl JJ and Reiser G. Characteristics of thrombin-induced calcium
signals in rat astrocytes. Glia 21: 361–369, 1997.
289. Ubl JJ, Sergeeva M, and Reiser G. Desensitisation of proteaseactivated receptor-1 (PAR-1) in rat astrocytes: evidence for a novel
mechanism for terminating Ca2⫹ signalling evoked by the tethered
ligand. J Physiol 525: 319 –330, 2000.
290. Ubl JJ, Vohringer C, and Reiser G. Co-existence of two types of
[Ca2⫹]i-inducing protease-activated receptors (PAR-1 and PAR-2)
in rat astrocytes and C6 glioma cells. Neuroscience 86: 597– 609,
1998.
291. Uehara A, Muramoto K, Takada H, and Sugawara S. Neutrophil
serine proteinases activate human nonepithelial cells to produce
inflammatory cytokines through protease-activated receptor 2.
J Immunol 170: 5690 –5696, 2003.
292. Uehara A, Sugawara S, Muramoto K, and Takada H. Activation
of human oral epithelial cells by neutrophil proteinase 3 through
protease-activated receptor-2. J Immunol 169: 4594 – 4603, 2002.
293. Vaughan PJ, Pike CJ, Cotman CW, and Cunningham DD.
Thrombin receptor activation protects neurons and astrocytes
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
257. Scott G, Deng A, Rodriguez-Burford C, Seiberg M, Han R,
Babiarz L, Grizzle W, Bell W, and Pentland A. Protease-activated receptor 2, a receptor involved in melanosome transfer, is
upregulated in human skin by ultraviolet irradiation. J Invest Dermatol 117: 1412–1420, 2001.
258. Scudamore CL, Thornton EM, McMillan L, Newlands GF, and
Miller HR. Release of the mucosal mast cell granule chymase, rat
mast cell protease-II, during anaphylaxis is associated with the
rapid development of paracellular permeability to macromolecules
in rat jejunum. J Exp Med 182: 1871–1881, 1995.
259. Seiberg M, Paine C, Sharlow E, Andrade-Gordon P, Costanzo
M, Eisinger M, and Shapiro SS. The protease-activated receptor
2 regulates pigmentation via keratinocyte-melanocyte interactions.
Exp Cell Res 254: 25–32, 2000.
260. Seiler SM, Peluso M, Michel IM, Goldenberg H, Fenton JWN,
Riexinger D, and Natarajan S. 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.
261. Shapiro MJ and Coughlin SR. Separate signals for agonist-independent and agonist-triggered trafficking of protease-activated receptor 1. J Biol Chem 273: 29009 –29014, 1998.
262. Shapiro MJ, Trejo J, Zeng D, and Coughlin SR. 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.
263. Shapiro MJ, Weiss EJ, Faruqi TR, and Coughlin SR. Proteaseactivated receptors 1 and 4 are shutoff with distinct kinetics after
activation by thrombin. J Biol Chem 275: 25216 –25221, 2000.
264. Sharlow ER, Paine CS, Babiarz L, Eisinger M, Shapiro S, and
Seiberg M. The protease-activated receptor-2 upregulates keratinocyte phagocytosis. J Cell Sci 113: 3093–3101, 2000.
265. Smith-Swintosky VL, Cheo-Isaacs CT, D’Andrea MR, Santulli
RJ, Darrow AL, and Andrade-Gordon P. Protease-activated receptor-2 (PAR-2) is present in the rat hippocampus and is associated with neurodegeneration. J Neurochem 69: 1890 –1896, 1997.
266. Southan C. A genomic perspective on human proteases. FEBS
Lett 498: 214 –218, 2001.
267. Stead RH, Tomioka M, Quinonez G, Simon GT, Felten SY, and
Bienenstock J. Intestinal mucosal mast cells in normal and nematode-infected rat intestines are in intimate contact with peptidergic nerves. Proc Natl Acad Sci USA 84: 2975–2979, 1987.
268. Steinhoff M, Corvera CU, Thoma MS, Kong W, McAlpine BE,
Caughey GH, Ansel JC, and Bunnett NW. Proteinase-activated
receptor-2 in human skin: tissue distribution and activation of
keratinocytes by mast cell tryptase. Exp Dermatol 8: 282–294, 1999.
269. Steinhoff M, Neisius U, Ikoma A, Fartasch M, Heyer G, Skov
PS, Luger TA, and Schmelz M. Proteinase-activated receptor-2
mediates itch: a novel pathway for pruritus in human skin. J Neurosci. In press.
270. Steinhoff M, Vergnolle N, Young SH, Tognetto M, Amadesi S,
Ennes HS, Trevisani M, Hollenberg MD, Wallace JL, Caughey
GH, Mitchell SE, Williams LM, Geppetti P, Mayer EA, and
Bunnett NW. Agonists of proteinase-activated receptor 2 induce
inflammation by a neurogenic mechanism. Nat Med 6: 151–158,
2000.
271. Striggow F, Riek M, Breder J, Henrich-Noack P, Reymann KG,
and Reiser G. The protease thrombin is an endogenous mediator
of hippocampal neuroprotection against ischemia at low concentrations but causes degeneration at high concentrations. Proc Natl
Acad Sci USA 97: 2264 –2269, 2000.
272. Suidan HS, Bouvier J, Schaerer E, Stone SR, Monard D, and
Tschopp J. 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.
273. Sun G, Stacey MA, Schmidt M, Mori L, and Mattoli S. Interaction of mite allergens der p3 and der p9 with protease-activated
receptor-2 expressed by lung epithelial cells. J Immunol 167: 1014 –
1021, 2001.
274. Suo Z, Wu M, Citron BA, Gao C, and Festoff BW. Persistent
protease-activated receptor 4 signaling mediates thrombin-induced
microglial activation. J Biol Chem. In press.
PROTEASE-ACTIVATED RECEPTORS
294.
295.
296.
297.
298.
300.
301.
302.
303.
304.
305.
306.
Physiol Rev • VOL
307.
308.
309.
310.
311.
312.
313.
314.
315.
316.
317.
318.
319.
Domains specifying thrombin-receptor interaction. Nature 353:
674 – 677, 1991.
Wakita H, Furukawa F, and MT. Thrombin and trypsin stimulate
granulocyte-macrophage colony-stimulating factor and interleukin-6 gene expression in cultured normal human keratinocytes.
Proc Assoc Am Phys 109: 190 –207, 1997.
Wang H and Reiser G. Thrombin signaling in the brain: the role of
protease-activated receptors. J Biol Chem 384: 193–202, 2003.
Wang H, Ubl JJ, and Reiser G. Four subtypes of proteaseactivated receptors, co-expressed in rat astrocytes, evoke different
physiological signaling. Glia 37: 53– 63, 2002.
Wang H, Ubl JJ, Stricker R, and Reiser G. Thrombin (PAR-1)induced proliferation in astrocytes via MAPK involves multiple
signaling pathways. Am J Physiol Cell Physiol 283: C1351–C1364,
2002.
Wang Y, Zhou Y, Szabo K, Haft CR, and Trejo J. Down-regulation of protease-activated receptor-1 is regulated by sorting nexin
1. Mol Biol Cell 13: 1965–1976, 2002.
Weinstein JR, Gold SJ, Cunningham DD, and Gall CM. 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.
Weiss EJ, Hamilton JR, Lease KE, and Coughlin SR. Protection against thrombosis in mice lacking PAR3. Blood 100: 3240 –
3244, 2002.
Widmann C, Gibson S, Jarpe MB, and Johnson GL. Mitogenactivated protein kinase: conservation of a three-kinase module
from yeast to human. Physiol Rev 79: 143–180, 1999.
Winitz S, Gupta SK, Qian NX, Heasley LE, Nemenoff RA, and
Johnson GL. Expression of a mutant Gi2 alpha subunit inhibits
ATP and thrombin stimulation of cytoplasmic phospholipase A2mediated arachidonic acid release independent of Ca2⫹ and mitogen-activated protein kinase regulation. J Biol Chem 269: 1889 –
1895, 1994.
Wojtukiewicz MZ, Tang DG, Ben-Josef E, Renaud C, Walz DA,
and Honn KV. Solid tumor cells express functional “tethered
ligand” thrombin receptor. Cancer Res 55: 698 –704, 1995.
Xu WF, Andersen H, Whitmore TE, Presnell SR, Yee DP,
Ching A, Gilbert T, Davie EW, and Foster DC. Cloning and
characterization of human protease-activated receptor 4. Proc Natl
Acad Sci USA 95: 6642– 6646, 1998.
Yang PC, Berin MC, Yu L, and Perdue MH. Mucosal pathophysiology and inflammatory changes in the late phase of the intestinal
allergic reaction in the rat. Am J Pathol 158: 681– 690, 2001.
Yu Z, Ahmad S, Schwartz JL, Banville D, and Shen SH. Proteintyrosine phosphatase SHP2 is positively linked to proteinase-activated receptor 2-mediated mitogenic pathway. J Biol Chem 272:
7519 –7524, 1997.
84 • APRIL 2004 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.33.6 on July 4, 2017
299.
from cell death produced by environmental insults. J Neurosci 15:
5389 –5401, 1995.
Vergnolle N. Proteinase-activated receptor-2-activating peptides
induce leukocyte rolling, adhesion, and extravasation in vivo. J Immunol 163: 5064 –5069, 1999.
Vergnolle N. Review article: proteinase-activated receptors: novel
signals for gastrointestinal pathophysiology. Aliment Pharmacol
Ther 14: 257–266, 2000.
Vergnolle N, Bunnett NW, Sharkey KA, Brussee V, Compton
SJ, Grady EF, Cirino G, Gerard N, Basbaum AI, AndradeGordon P, Hollenberg MD, and Wallace JL. Proteinase-activated receptor-2 and hyperalgesia: a novel pain pathway. Nat Med
7: 821– 826, 2001.
Vergnolle N, Derian CK, D’Andrea MR, Steinhoff M, and Andrade-Gordon P. Characterization of thrombin-induced leukocyte
rolling and adherence: a potential proinflammatory role for proteinase-activated receptor-4. J Immunol 169: 1467–1473, 2002.
Vergnolle N, Hollenberg MD, Sharkey KA, and Wallace JL.
Characterization of the inflammatory response to proteinase-activated receptor-2 (PAR2)-activating peptides in the rat paw. Br J
Pharmacol 127: 1083–1090, 1999.
Vergnolle N, Hollenberg MD, and Wallace JL. Pro- and antiinflammatory actions of thrombin: a distinct role for proteinaseactivated receptor-1 (PAR1). Br J Pharmacol 126: 1262–1268, 1999.
Vergnolle N, MacNaughton WK, Al-Ani B, Saifeddine M, Wallace JL, and Hollenberg MD. Proteinase-activated receptor 2
(PAR2)-activating peptides: identification of a receptor distinct
from PAR2 that regulates intestinal transport. Proc Natl Acad Sci
USA 95: 7766 –7771, 1998.
Vogel SM, Gao X, Mehta D, Ye RD, John TA, Andrade-Gordon
P, Tiruppathi C, and Malik AB. Abrogation of thrombin-induced
increase in pulmonary microvascular permeability in PAR-1 knockout mice. Physiol Genomics 4: 137–145, 2000.
Vouret-Craviari V, Auberger P, Pouyssegur J, and Van Obberghen-Schilling E. Distinct mechanisms regulate 5-HT2 and
thrombin receptor desensitization. J Biol Chem 270: 4813– 4821,
1995.
Vouret-Craviari V, Boquet P, Pouyssegur J, and Van Obberghen-Schilling E. Regulation of the actin cytoskeleton by
thrombin in human endothelial cells: role of Rho proteins in endothelial barrier function. Mol Biol Cell 9: 2639 –2653, 1998.
Vouret-Craviari V, Grall D, Chambard JC, Rasmussen UB,
Pouyssegur J, and Van Obberghen-Schilling E. Post-translational and activation-dependent modifications of the G proteincoupled thrombin receptor. J Biol Chem 270: 8367– 8372, 1995.
Vu TK, Hung DT, Wheaton VI, and Coughlin SR. Molecular
cloning of a functional thrombin receptor reveals a novel proteolytic mechanism of receptor activation. Cell 64: 1057–1068, 1991.
Vu TK, Wheaton VI, Hung DT, Charo I, and Coughlin SR.
621