Vous êtes sur la page 1sur 68

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

1 de 68

Editors:

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

Colman, Robert W.; Clowes, Alexander W.;

Goldhaber, Samuel Z.; Marder, Victor J.; George, James N.


Title:

Hemostasis and Thrombosis: Basic Principles and

Clinical Practice, 5th Edition


Copyright 2006 Lippincott Williams & Wilkins
> Tabl e of C ontents > P art I - B asi c Pri nci pl es of Hem ostasi s and
Throm bosi s > Secti on B - Fi bri nol ysi s and i ts Regul ati on > C hapter 19 Pl as m i nogen Acti vator I nhi bi tor-1

Chapter 19
Plasminogen Activator Inhibitor-1
Manuel Yepes
David J. Loskutoff
Daniel A. Lawrence
Plasminogen activators (PAs) are specific serine proteinases that
activate the proenzyme plasminogen, by cleavage of a single
Arg-Val peptide bond, to the broad-specificity enzyme plasmin.
Plasminogen activation provides an important source of localized
proteolytic activity during a number of physiologic and pathologic
processes such as fibrinolysis, ovulation, cell migration, epithelial
differentiation, vascular disease, and cancer (1,2,3). Therefore,
accurate regulation of PAs constitutes a critical feature of many
physiologic and pathologic events. Two PAs are found in mammals:
tissue-type PA (tPA) and urokinase-type PA (uPA) (4). The
regulation of PAs is a complex process that involves regulation of
gene expression by factors such as hormones, growth factors, and
cytokines (1,4), as well as regulation of enzyme activity through
interactions with fibrin (5) or with specific receptors for uPA (6),
tPA (7), and plasminogen (8). PA activity is also regulated by
specific inhibitors termed plasminogen activator inhibitors (PAIs)
(9). However, of the five different PAIs [i.e., PAI-1 (10); PAI-2
(11); PAI-3, also called the activated protein C inhibitor (APCI)
(12); protease nexin-1 (13); and neuroserpin (14)], only PAI-1,
which is the most kinetically efficient, appears to play a significant

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

2 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

role in regulating PA activity in blood and most tissues. The


exception to this may be in the central nervous system (CNS),
where neuroserpin is an important regulator of tPA activity
(15,16). This chapter focuses on the biochemical, genetic,
physiologic, and pathologic properties of PAI-1. The basic
properties of PAI-1 are analyzed in the first part, whereas the
second half examines the evidence linking PAI-1 to the
pathogenesis of human disease.

BASIC PROPERTIES OF PLASMINOGEN


ACTIVATOR INHIBITOR-1
Protein Structure and Function
Formerly known as the endothelial cell PAI (10,17,18,19), the
fast-acting inhibitor of tPA in plasma (20), and the PAI
(21,22,23), PAI-1 is a single-chain glycoprotein having an M r of
approximately 50,000 and was first identified in the year 1983
(10,24,25). PAI-1 has three potential N-linked glycosylation sites
and contains between 15% and 20% carbohydrate (19,23). It
belongs to the serine proteinase inhibitors superfamily (serpins),
which is a gene family that includes many of the proteinase
inhibitors found in blood, as well as other proteins with unrelated
or unknown functions (26) (see Chapter 13). The serpins share a
tertiary structure (27) and have evolved from a common ancestor
(28,29). They act as suicide inhibitors [i.e., they react only once
with their target protease to form sodium dodecyl sulfate
(SDS)stable complexes (30,31,32,33)]. The interaction between
the serpin and its target protease occurs at an amino acid residue
located in the reactive center loop (RCL) of the serpin, which is
known as a bait residue. This bait residue is thought to mimic
the normal substrate of the enzyme and to associate by its
side-chain atoms with the specificity crevice (S1 site) of the
enzyme (34,35,36,37). The bait amino acid is called the P1
residue, with the amino acids toward the amino-terminal side of
the scissile RCL bond being labeled in the order P1, P2, P3, and so
on, and the amino acids on the carboxyl side being labeled P1,
P2, P3, and so on (38). Upon cleavage of the bait amino acid by
a target proteinase, there is a large conformational change in the
serpin that involves the rapid insertion of the RCL into the major

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

3 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

structural element of the serpin, the -sheet A (33,39,40,41). This


results in a tight docking of the enzyme to the serpin surface and
in a large increase in the structural stability of the serpin. This
leads to distortion of the enzyme structure, including the active
site, which traps the proteinase in an acylenzyme complex with
the serpin (42,43,44).
The PAI-1 complementary deoxyribonucleic acid (cDNA) encodes a
protein of 402 amino acids that includes a typical secretion signal
sequence (23,45,46,47). Mature human PAI-1 is composed of two
variants in approximately equal proportions (i.e., 381 and 379
amino acids), which likely arise from alternative cleavage of the
secretion signal sequence and generate proteins with overlapping
amino-terminal sequences of Ser-Ala-Val-His-His and
Val-His-His-Pro-Pro (45,48). The latter sequence does not contain
cysteines because the single cysteine residue present in the signal
peptide is removed during membrane translocation (46,47). This
property facilitates the efficient expression and isolation of
recombinant PAI-1 from Escherichia coli (48,49,50,51,52), which,
in contrast to PAI-1 purified from mammalian cell culture, is
predominantly produced in the active form (48).

Reactive Center Loop


The P1 bait amino acid of PAI-1 (Arg346) is contained within the
RCL near the carboxyl terminus of the molecule and serves as a
pseudosubstrate for the target serine proteinase (33,37). Either
arginine or lysine at P1 is essential for PAI-1 to function as an
effective inhibitor of uPA (36,37), and the residues surrounding P1
can modulate PAI-1 inhibitory activity by up to two orders of
magnitude and can also alter the targetprotease specificity
(42,53,54). In the case of tPA other P1 residues are tolerated, and
this is most likely due to tighter exosite interactions between
PAI-1 and tPA (53,55,56,57).

Plasminogen Activator Inhibitor-1


Conformations
Native PAI-1 exists in at least two distinct conformations: an
active form that is produced by cells and is secreted and an
inactive

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

4 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

P.366
or latent form that accumulates in cell culture medium over time
(58,59,60) (see Fig. 19-1). In blood and tissues, most of the PAI-1
is in the active form; however, in platelets both the active and the
latent PAI-1 are found (21). In the active form of PAI-1, the RCL is
part of an exposed loop on the surface of the molecule (61). Upon
reaction with a proteinase, this RCL cleaves and integrates into the
center of its own -sheet A (33,61). In the latent form, the RCL is
intact, but instead of being exposed, the entire amino-terminal
side of the RCL is inserted as the central strand into the -sheet A
(60). This insertion accounts for the increased stability of latent
PAI as well as for its lack of inhibitory activity (58,62,63). The
active form spontaneously converts to the latent form with a
half-life of about 1 to 2 hours at 37C at neutral or slightly
alkaline pH (48,59,64,65,66). The latent form can also be
converted into the active form by treatment with denaturants or
negatively charged phospholipids, or can be converted very slowly
in the presence of the protein vitronectin (58,66,67). This
spontaneous reversible interconversion between the active and
latent structures is unique for PAI-1 and distinguishes it from other
serpins.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

5 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

FIGURE 19-1. Schematic illustration of the three


conformations of plasminogen activator inhibitor-1 (PAI-1). The
major -sheet (-sheet A) is highlighted in light gray, and the
reactive center loop (RCL) is in dark gray. A: The active
conformation of a stable mutant of PAI-1 (61). B: The latent
conformation of PAI-1 (60). C: The cleaved conformation of
PAI-1 (287). In both the latent and cleaved forms of PAI-1, the
RCL is inserted into -sheet A to form a new -sheet (strand
4A) (see Color Fig. 19-1). (From Sharp AM, Stein PE, Pannu
NS, et al. The active conformation of plasminogen activator
inhibitor 1, a target for drugs to control fibrinolysis and cell
adhesion. Structure Fold Des 1999;7:111118, with
permission.)

Other inactive forms of PAI-1 have also been identified. The first
form results from the oxidation of one or more critical methionine
residues within active PAI-1 (68,69). This form differs from latent
PAI-1 in that it can be partially reactivated by treatment with an
enzyme that specifically reduces oxidized methionine residues
(68). Oxidative inactivation of PAI-1 may be an additional
mechanism for the regulation of the PA system. Oxygen radicals
produced locally by neutrophils or other cells could inactivate
PAI-1 and thereby facilitate the generation of plasmin activity at
sites of infections or in areas of tissue remodeling (70). A fourth
conformational form of PAI-1 has also been identified. This is a
noninhibitory substrate form that can be induced by the addition of
SDS and can be converted back to the active or latent forms by
treatment with 4 M guanidine HCl (71,72). The significance of this
form and whether it exists in vivo is not known. PAI-1 can also
exist in two different cleaved forms. As noted previously, PAI-1
that is complexed with a proteinase is cleaved at the site, and
PAI-1 can also be found not in complex with a proteinase but with
its RCL cleaved. This can arise from dissociation of the PAI-1PA
complex or from the cleavage of the RCL by a nontarget proteinase
at a site other than the (63,73). None of these forms of PAI-1 is
able to inhibit proteinase activity; however, they may interact with
other nonproteinase substrates.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

6 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

Biochemical Properties
Interaction with Tissue-Type Plasminogen
Activator, Urokinase-Type Plasminogen
Activator, and Plasmin
Inhibition of PAs by PAI-1 occurs in a rapid and stoichiometric
manner, resulting in the formation of a covalent bond between the
two molecules. Several studies indicate that PAI-1 is cleaved
during this reaction and that the amino-terminal end of the RCL
inserts as an antiparallel strand into -sheet A (74,75). The
inhibitor is consumed in the process, giving rise to the discussed
term suicide inhibitor.
P.367
Kinetic studies demonstrate that PAI-1 inhibits the naturally
occurring single-chain form of tPA with a second-order rate
constant of approximately 10 6 M - 1 S -1 , a value that is at least
1,000 times higher than those for the interactions of PAI-2, PAI-3,
and protease nexin-1 with single-chain tPA (9,42). Moreover,
approximately 70% of the total tPA in carefully collected normal
human plasma is detected in complex with PAI-1, suggesting that
the inhibition of tPA by PAI-1 is a normal, ongoing process. PAI-1
is also an important uPA inhibitor because the second-order rate
constant for its interaction with uPA is also at least two orders of
magnitude higher than that of other PAIs (9,42,76). PAI-1 can also
efficiently and directly inhibit plasmin (64,77). Therefore, PAI-1 is
the chief regulator of plasmin generation in vivo, and, as such, it
appears to play an important role in both fibrinolytic and
thrombotic diseases (3,78,79,80).

Interaction with Vitronectin and Members of


the Low Density Lipoprotein Receptor Family
(LDL-Rs)
Most of the active PAI-1 in blood circulates in the form of a
complex with the glycoprotein vitronectin. The binding site of
vitronectin to PAI-1 has been identified in a region centered
around residues 101 to 123 of the three-dimensional structure
(81,82,83,84), whereas the binding site to members of the LDL-R

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

7 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

family has been less well characterized and appears to be located


in a region associated with helix D containing residues Arg76 and
Lys69 (85,86). Vitronectin is present in plasma and in the
extracellular matrix (87,88), mainly at sites of injury or
remodeling. Vitronectin is also specifically incorporated into fibrin
clots (89). Vitronectin may be considered a cofactor for PAI-1
because it stabilizes PAI-1 in its active conformation, thereby
increasing its biologic half-life. In turn, PAI-1 converts vitronectin
from its native, plasma form, which does not support cell adhesion,
to an activated form that is able to bind ligands such as integrins
(90). Vitronectin also increases the inhibitory efficiency of PAI-1
for thrombin approximately 300-fold, making it a more efficient
inhibitor of thrombin than antithrombin III in the absence of
heparin (91,92) (see Fig. 19-2).
Upon complex formation with a proteinase, the conformational
change in PAI-1 associated with RCL insertion results in a loss in
high affinity binding to vitronectin but in a gain in high affinity
binding to the clearance receptors of the (LDL-R) family (85,93).
This is a result of a conformational change in PAI-1 that disrupts
the vitronectin binding site, exposing, at the same time, a cryptic
receptor binding site that is revealed only when PAI-1 is in an
active conformation complex with a proteinase (54,61,85). This
results in an approximately 1,000,000-fold shift in the relative
affinity of PAI-1 from vitronectin to a member of the LDL-R family,
which leads to a rapid clearance of PAI-1 by internalization through
the LDL-R family members (85).
The uPA receptor (uPAR) (6) is a glycosyl phosphatidyl inositol
(GPI)anchored protein that localizes uPA on the cell surface,
frequently at the invading edge of cells (94). Receptor occupancy
has been shown to activate intracellular signaling pathways. Like
PAI-1, uPAR also binds to vitronectin with high affinity as well as
to various integrins (95,96,97,98), and the binding of uPAR to
vitronectin is inhibited by PAI-1. The fact that uPAR and integrins
also may bind to vitronectin raises the possibility that, as
discussed later in this chapter, PAI-1 could regulate cell adhesion
and migration.

Binding to Heparin and Fibrin

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

8 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

PAI-1 also binds to heparin with high affinity (48,99). This binding
does not affect the interaction of PAI-1 with uPA or tPA but, in
contrast, enhances the interaction of PAI-1 with thrombin
(91,100). The heparin binding domain on PAI-1 has been mapped
to a region homologous to the heparin binding domain of
antithrombin III on and around helix D (101). Critical residues
appear to include lysines 65, 69, 80, and 88, and arginine 76. It
has also been reported that PAI-1 binds to fibrin in vitro with a K d
of 3.8 M, and, while bound, remains capable of inhibiting uPA and
tPA (102,103,104,105). However, more recent data suggest that
most of the PAI-1 localized to fibrin clots is vitronectin dependent
(106,107).

Structure and Regulation of the


Plasminogen Activator Inhibitor-1 Gene
The human gene for PAI-1 is located on chromosome 7q21.3-22
(108). It is 12.3 kb in length, composed of 9 exons and 8 introns,
and is similar in structure to that of protease nexin I and
neuroserpin (109,110,111,112,113,114,115). The promoter
contains a typical TATA box but no CAAT sequence. Segments of
PAI-1 promoter containing as little as 187 bp of upstream
sequence have shown to direct transcription in several mammalian
cell types (111,116,117). Comparison of the rat and human PAI-1
promoter sequences shows a striking region of conservation in
P.368
the proximal promoter (from the TATA box to -90) and a distal
sequence of -510 to -753 of the rat sequence. Changes in plasma
PAI-1 levels can be correlated with variations in the structure of
the PAI-1 gene. To date, three polymorphic variations in the gene
have been reported: a single nucleotide insertion per deletion
(4G/5G) polymorphism in the promoter region, an allelic variation
at a (C-A) n dinucleotide repeat polymorphism in intron 3, and a
HindIII restriction fragment length polymorphism (RFLP) due to a
base change at the 3 flanking region of the gene (118). However,
no link between PAI-1 polymorphism and human disease has been
consistently demonstrated. Relative to the HindIII RFLP
polymorphism, three genotypes1/1, 1/2, and 2/2have been
described, on the basis of the presence (2 allele) or the absence

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

9 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

(1 allele) of the HindIII polymorphic site. Interestingly, genotype


1/1 exhibits higher plasma PAI-1 activity than genotype 2/2 (119).
Two sizes of PAI-1 messenger ribonucleic acid (mRNA) are
observed in human cells, approximately 3 kb and 2 kb in length,
respectively. This difference has been shown to be because of
alternative polyadenylation, with both mRNAs encoding the same
protein, but with an additional 1 kb of 3 untranslated region
present in the larger message (109,110,111,112). The additional
3 untranslated region in the larger message contains a 75-base
pair AT-rich sequence that has been postulated to play a role in
PAI-1 gene regulation by a posttranscriptional mechanism
(47,120,121,122,123).

FIGURE 19-2. Plasminogen activator inhibitor-1 (PAI-1)


association with vitronectin and low density lipoprotein
receptors (LDL-Rs) is conformationally controlled. Surface
plasmin resonance analysis of PAI-1 or PAI-1uPA complexes
binding to either vitronectin or the LDL-R family member LRP.
The solid lines in each panel are PAI-1 only and the dashed
lines in each panel are the PAI-1uPA covalent complex. (From
Natalia Gorlatova and Daniel A. Lawrence.)

Although PAI-1 is present at low concentrations in plasma, its


relatively short half-life in blood (i.e., 10 minutes) suggests a high
biosynthetic rate (124). Moreover, its concentration rapidly
increases in response to a variety of agents or changes in

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

10 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

physiologic state, indicating that the amount of PAI-1 in plasma is


subject to dynamic regulation. For example, the concentration of
plasma PAI-1 increases dramatically during endotoxemia (125).
Endotoxin [lipopolysaccharide (LPS)] induces PAI-1 mRNA in
virtually all tissues of the mouse (126), suggesting that plasma
PAI-1 may originate from multiple tissues during sepsis. In situ
hybridization analysis reveals that endotoxin induces PAI-1 mRNA
in endothelial cells at all levels of the vasculature, including larger
arteries, veins, and capillaries. PAI-1 gene expression is also
induced in hepatocytes and in adipocytes (127,128).
Many of the effects of endotoxin are mediated through the release
of cytokines from inflammatory cells [e.g., tumor necrosis factor-
(TNF-) and interleukin-1 (IL-1)]. Indeed, a variety of studies
have shown that most of the same tissues of the mouse that
produce PAI-1 in response to endotoxin (i.e., liver, heart, and
lung) also produce it in response to TNF- (126) Therefore,
endotoxin and TNF- upregulate PAI-1 expression primarily in
endothelial cells in most mouse tissues and induces it in vitro in
various bovine and human endothelial cells. Growth modulators
such as transforming growth factor- (TGF-) also induce plasma
PAI-1 antigen and tissue PAI-1 mRNA in several animal models
(126). In murine models, TGF- induces PAI-1 mRNA in vascular
and nonvascular smooth muscle cells, in adipocytes, and in cells in
the myocardium and kidney (126). Although TGF- induces PAI-1
in cultured bovine endothelial cells (129), it does not appear to
induce it in murine endothelium in vivo. Whether this apparent
inconsistency reflects differences between in vitro and in vivo
systems or between species remains to be determined.
Several studies have identified DNA sequence elements in the first
1.3 kb of the promoter/5-upstream flanking region of the PAI-1
gene that mediate cytokine responsiveness in transfected cells
(130). However, little is known about the role of these sequences
in PAI-1 promoter function in vivo. PAI-1 mRNA levels are
determined by both transcriptional and posttranscriptional
mechanisms. Factors that have been shown to increase the rate of
PAI-1 mRNA transcription in different cell types are
glucocorticoids, TNF- insulin, and IL-1 (130). In HepG2 cells,
IGF-1 (insulinlike growth factor-1) and insulin are also able to

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

11 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

induce PAI-1 synthesis by increasing mRNA stability (131).


Moreover, it has been suggested that PAI-1 gene regulation is of
even greater complexity because it also could depend on genetic
polymorphisms.
Expression of PAI-1 has been observed in a wide range of cell
types including adipocytes, fibrosarcoma, hepatoma, and ovarian
and endothelial cells. However, studies in transgenic mice carrying
the PAI-1 promoter linked to an indicator gene suggest that PAI-1
expression may be much more limited in vivo (126). Although the
hepatoma cells produce large amounts of PAI-1 in vitro and have
been used as a model to study its regulation, it is known that
hepatocytes do not normally synthesize PAI-1 in vivo (132,133).
Nevertheless, these hepatocytes and the endothelial cells can be
induced by endotoxin to produce PAI-1 in vivo (133). These
observations raise important concerns about the validity of in vitro
tissue culture as a model for the regulation of fibrinolysis in vivo.

PLASMINOGEN ACTIVATOR INHIBITOR-1


IN PATHOLOGIC CONDITIONS
In healthy individuals, PAI-1 is expressed primarily in
megakaryocytes, smooth muscle cells, and adipocytes (128,
130,134,135). However, as noted in the preceding text, its
expression can be rapidly and markedly induced in many cell types
by stress, or injury, or by growth factors and cytokines
(133,136,137). The last part of this chapter analyzes the link
between PAI-1 and different pathologic events such as cancer,
obesity, atherosclerosis, and vascular, pulmonary, and renal
diseases (see Fig. 19-3).

Cancer
The existence of a link between fibrinolysis and tumor growth was
first suggested by the observations that patients with malignant
tumors have increased fibrinolytic activity and that tumor tissues
can degrade fibrin clots (1,138,139,140). Therefore, it was quite
unexpected when high PAI-1 levels were found to be strongly
correlated with a poor prognosis in patients with many different
neoplasias, including gastric and breast carcinomas, as well as with
brain, ovarian, and lung tumors and with metastatic lesions from

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

12 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

renal cell carcinoma, melanoma,


P.369
and colorectal cancer (141,142,143,144,145,146). Analysis of
human colon carcinomas has demonstrated the presence of PAI-1
mRNA in vascular endothelial cells, in the stroma surrounding the
invasive tumor, in granulation tissue, and in some capillaries
within the tumor (147). Strong staining for PAI-1 antigen has also
been observed in proliferative vessels in intracranial tumors such
as high-grade gliomas and metastatic tumors, as well as in blood
vessels near the necrotic center of tumors (147,148). In these
studies, PAI-1 was localized to the vascular basement membrane
and perivascular connective tissue, whereas endothelial cells
themselves showed only weak reactivity. In contrast, an
immunochemical study on Lewis lung carcinoma transplanted in
mice showed PAI-1 protein in the tumor cells themselves (149).
Like PAI-1, uPA and its receptor, uPAR, have also been correlated
with a poor prognosis in cancer (150,151). In this section of the
chapter, the role of PAI-1 in cancer has been addressed from two
perspectives: cell migration and tumor angiogenesis.

FIGURE 19-3. Plasminogen activator inhibitor-1 (PAI-1)

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

13 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

association with normal and pathologic physiology.

Cell Migration
Although uPA has been associated with physiologic events that
involve cell migration such as wound healing, vascular remodeling
(152,153,154), neural cell migration (155), and monocyte
invasiveness (156), its precise role in these processes is not clear.
Increased uPA levels have been reported in many transformed cells
including myeloid leukemic cells (157), hepatomas (158), gliomas
(159), and carcinomas (160), and inhibitors of uPA have been
shown to reduce cell migration and metastasis (161,162,163,164).
The long-established explanation of these data was that the main
role of uPA was to activate plasmin, which directly, and indirectly
through the activation of matrix metalloproteases (MMPs),
degrades the basement membrane and clears a path for cells to
migrate (165,166) (see Fig. 19-4A). However, by the 1990s, this
matrix barrier model of proteases and cell migration started to
be challenged. For example, it was shown that uPA and plasmin as
well as other proteases could produce a limited proteolysis that
preserved the architecture of the basement membrane but exposed
cryptic sites within the matrix that could enhance cell adhesion and
or migration (167,168,169,170,171). Plasmin could also promote
cell proliferation and migration through the activation of growth
factors such as TGF- (172,173) or through the release of
sequestered growth factors from the matrix such as fibroblast
growth factor (FGF) and vascular endothelial growth factor (VEGF)
(174,175,176,177). More recent studies have expanded on these
observations and have begun to suggest that the role of uPA and,
in particular, PAI-1 in cell migration may be far more subtle and
elegant than previously thought and may involve interactions on
the cell surface with a number of receptors including uPAR,
integrins, and members of the LDL-R family.
Cell surface uPA and uPAR were originally shown to localize at the
focal contacts (178,179) and at the leading edge of migrating cells
(180). As already noted, uPAR not only serves as a receptor for
uPA but also binds to vitronectin in the matrix, and, through this

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

14 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

binding, can mediate cell adhesion directly (181). uPAR can also
associate either directly or indirectly with integrins, and this
association can affect both cell signaling and migration
(95,96,97,98,182,183). In vitro studies have demonstrated that
both PAI-1 and uPAR bind to the somatomedin B-domain of
vitronectin, although uPAR binds with significantly lower affinity
(184,185). Therefore, PAI-1 competes with uPAR for binding to
vitronectin and may inhibit uPAR-mediated cell attachment
(185,186,187). PAI-1, uPA, and uPAR can also modulate
integrin-mediated cell adhesion (167,168,188). For example, the
integrin v 3 binding site on vitronectin overlaps the PAI-1 binding
site, resulting in a potential competition between v 3 and PAI-1
for vitronectin binding (167,184). Therefore, in smooth muscle
cells where vitronectin promotes cell migration, the inhibition of
v 3 binding to vitronectin by the addition of PAI-1 results in an
inhibition of cell migration (167). Exogenous PAI-1 also inhibits the
migration of stimulated endothelial, human amnion WISH and
epidermal carcinoma cells, as well as the invasion of human
monocytes (167,168,189,190). However, the effects of PAI-1 on
cell adhesion and migration are fully reversible by PAs because, as
discussed at the beginning of this chapter, upon complex formation
with a protease, PAI-1 undergoes a conformational change that
alters the PAI-1 vitronectin binding site and renders the PAPAI-1
complex unable to bind to vitronectin. Therefore, the relative
concentrations of uPA (or tPA) and active PAI-1 at the cell matrix
interface are able to regulate cell-matrix interactions mediated
through vitronectin.
In addition to blocking cell-matrix interactions and cell migration,
PAI-1 can also promote cell migration under some conditions. In
the studies discussed in the preceding text, high concentrations of
exogenous PAI-1 were used to block migration. However, recent
work by Palmieri et al. has established that the stable expression
of PAI-1 can actually stimulate adhesion and migration on several
different matrix proteins (191). And although this activity required
the inhibition of uPA, it did not specifically require PAI-1, because
another uPA inhibitor, PAI-3, could also stimulate this activity. The
inhibition of uPA by either PAI-1 or PAI-3 also increased the
surface expression of several integrin subunits. Czekay et al.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

15 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

demonstrated that the binding of uPA to uPAR promotes the


association of the uPAuPAR complex with the integrins v 3 or
v 3 and showed that by adding PAI-1 to this complex, they could
induce the cells to detach from the matrix by stimulating
endocytosis of the integrinuPARuPAPAI-1 complex by a member
of the LDL-R family (188). The endocytosed integrins could then
recycle back to the cell surface, where, upon activation, they could
reengage the matrix. These studies have lead to the hypothesis
that the inhibition of uPA by PAI-1 could promote the type of
cycled attachmentdetachmentreattachment required for cell
migration (166,188). Importantly, this activity appears to be
independent of the matrix composition, suggesting that it may
represent a general mechanism in which inhibition of uPA promotes
cell migration (Fig. 19-4B). This model is consistent with much of
the literature examining the roles of uPA and PAI-1 in cell
migration. It provides a compelling explanation as to why both uPA
and its inhibitor PAI-1 are strongly correlated with a poor
prognosis in cancer patients because both molecules are needed
for the maximal enhancement of cell migration. In this context, it
is also interesting to note that PAI-1 differs from other serpins in
that it is a trace protein in plasma and tissues, with a relatively
short half-life (approximately 10 minutes). Moreover, it is an
immediate early gene and has been shown to accumulate at focal
points of adhesion (192). Finally, its biosynthesis is stimulated
rapidly by a variety of inflammatory mediators, growth factors, and
hormones (130,136). The short half-life and ability to be rapidly
and dramatically unregulated are the expected properties of
molecules with the potential to rapidly initiate or terminate biologic
processes (i.e., molecular switches) (80,185). Therefore, by
regulating the production rate or activity of PAI-1, cells may be
able to control their adhesiveness and movement.
Finally, PAI-1 can also regulate cell motility by effecting
uPA-dependent cell signaling by phosphorylation of the
extracellular signal-regulated kinase (ERK). This signaling requires
uPA binding to uPAR and requires endocytosis of the
PAI-1uPAuPAR complex by the VLDLR and recycling of uPAR back
to the cell surface (193). The effect of PAI-1 inhibition of
P.370

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

16 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

uPA and endocytosis of the complex is unclear: It was shown to


inhibit uPA-induced chemotaxis in one study (194), whereas in
another report, PAI-1uPA complexes increased cell migration and
proliferation in a process that required association with the LDL-R
family and was dependent upon uPAR recycling (193).

FIGURE 19-4. The evolution of our understanding of the role


of plasminogen activator inhibitor-1 (PAI-1) in cell migration.
A: Before the multiple actions of urokinase PA (uPA),
urokinase-type plasminogen activator receptor (uPAR), and

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

17 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

PAIs were fully understood, the simplest model for the function
of uPA in cell migration was based solely on its ability to
activate plasmin. Plasmin in turn degraded extracellular matrix
(ECM) proteins, which allowed the cell to escape its matrix
barrier and migrate. B: A model of the potential role of PAI-1
in cell migration that assumes that the cell surface receptors
uPAR, integrins, and low density lipoprotein receptor (LDL-R)
family, such as the very low density lipoprotein receptor
(VLDLR), are expressed by the cell. The integrin may or may
not be attached to the matrix and connected to the
cytoskeleton, and uPA may be expressed by the same cell or by
a nearby cell. Binding of the secreted uPA to uPAR may
enhance the association of uPAR with an integrin, which may
also promote adhesion. If there is also PAI in the surrounding
milieu, it can then bind to the uPA present in the ternary
complex. This PAI would most likely be PAI-1 because its
expression is stimulated by conditions associated with cell
migration and because it will specifically localize to the matrix
through binding to vitronectin. The binding of PAI-1 to uPA
induces the conformational change in PAI-1 that simultaneously
reduces its affinity for the matrix while enhancing its affinity
for the clearance receptor (Fig. 19-2). This association also
promotes integrin disengagement from the matrix, and very
likely from the cytoskeleton as well, and binding to the
clearance receptor (VLDLR). Whether the integrin first
disengages from the matrix and then binds to the clearance
receptor or whether it is the association of the quaternary
complex with the clearance receptor that induces the integrin
to disengage is not yet clear. Regardless of the exact sequence
of events, the quaternary complex is then endocytosed, and, in
the endosome, the PAIuPA complex separates from the three
receptors and is targeted to the lysosome for degradation. The
LDL-R, uPAR, and the integrin are then recycled back to the
cell surface where the process can be repeated. This cycled
attachmentdetachmentreattachment of integrins is necessary
for cell migration. This figure does not indicate the many

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

18 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

potential intracellular signaling events that each one of these


interactions could generate, but each one of these receptor
interactions could signal the cell. The downstream
consequences of these events undoubtedly also modulate cell
adhesion and migration through other pathways (see Color Fig.
19-4). (From Stefansson S, Lawrence DA. Old dogs and new
tricks, proteases, inhibitors, and cell migration. Sci STKE
2003;e24, with permission.)

Tumor Angiogenesis
Angiogenesis is an important factor for tumor growth and
metastasis. It has been proposed that extracellular matrix
remodeling is necessary to allow invasion of the newly forming
blood vessels, which would explain why uPA and its receptor
P.371
are elevated in several types of cancers. However, it is the spatial
relationship between uPA and PAI-1 that seems to be critical for
the tumorigenic response. Immunohistochemical staining and in
situ hybridization of aortic explants and cocultures of endothelial
cells and fibroblasts have shown uPA to be predominantly located
in the sprouting endothelium, whereas PAI-1 expression is strong
in the population of stromal fibroblasts that are directly in contact
with the migrating endothelial cells (195). These findings have also
been described in breast carcinomas, where PAI-1 expression is
found primarily in fibroblasts, and this expression is associated
with an increased incidence of tumor invasion (196,197).
The effects of PAI-1 on tumor growth and angiogenesis in ectopic
or transplant tumor models are variable and are sometimes
opposed (77,198,199,200,201,202,203,204,205,206). A potential
explanation for these differences is that PAI-1, in vivo, normally
plays a regulatory role to either enhance or reduce cell migration
and/or angiogenesis during wound healing. This system then may
also play a role in tumor growth, and the potential importance of
this regulation is likely to be very context specific. For example, in
transplant tumor models, small differences in the initial conditions
of either the tumor or the recipient of the transplant may have

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

19 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

very large effects on early events such as the number of


transplanted cells that survive or the adhesion and engraftment of
the tissue, and it may be these very early events that are largely
determining the final outcome of the experiments. Nevertheless,
the potential for PAI-1 to regulate angiogenesis has been shown in
several nontumor models, where PAI-1 has been found to be a
potent regulator of angiogenesis (204). The inhibitory effect of
PAI-1 on angiogenesis in this model can be explained by the ability
of PAI-1 to inhibit both integrin access to vitronectin and
proteinase activity (77).
A dual role for PAI-1 in tumor growth and angiogenesis has been
demonstrated by the finding that physiologic levels of PAI-1
promote the growth of M21 human melanoma tumors in nude mice,
whereas pharmacologic levels of PAI-1 are inhibitory (204). This
pattern of growth is mirrored in the extent of angiogenesis, with
the treatment of tumors with low doses of PAI-1 increasing the
density of vessel branching, whereas substantially fewer branches
are observed upon treatment with higher doses (80,204). This
dose-response curve has also been observed in Matrigel implants
in mice (204) (see Fig. 19-5) and in an ex vivo aortic ring explant
assay (205). Therefore, PAI-1 concentrations at or near to the
normal physiologic range appear to promote angiogenesis, whereas
pharmacologic levels of PAI-1 appear to inhibit angiogenesis.
These data yield a classic bell-shaped curve for the effects of
PAI-1 on angiogenesis, consistent with its potential role as a
regulator of angiogenesis in vivo.

Clinical Implications
As noted previously, high tumor levels of PAI-1 are one of the
most informative markers of poor prognosis in several types of
cancers (141,142,143,144,145,146). This would suggest that,
despite the various results obtained in different in vivo andin vitro
models of tumor growth and angiogenesis, PAI-1 might be used as
a prognostic marker in patients with cancer, as well as a predictive
factor for therapy response. Moreover, these results also suggest
that therapeutic modulation of PAI-1 levels might be a potential
target for anticancer treatment.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

20 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

Plasminogen Activator Inhibitor-1 in


Vascular Disease
In addition to tumor growth and angiogenesis, PAI-1 is also
important in vascular disease. In this section, the role of PAI-1 in
vascular disease is addressed from four different perspectives:
vascular fibrinolysis, vascular wound healing, atherosclerosis, and
cardiovascular and cerebrovascular disease.

Vascular Fibrinolysis
The coagulation cascade is a stepwise activation and amplification
of plasma proteinases resulting in the conversion of prothrombin to
thrombin, with the generation of the fibrin clot. The coagulation
cascade also leads to platelet activation
P.372
by a proteinase-activated receptor (207). Activated platelets then
release their granular content into the clot, which includes PAI-1
and vitronectin. The dissolution of the fibrin clot is primarily
mediated by tPA, which is synthesized and stored within
endothelial cells. tPA binds fibrin and is incorporated into the clot
along with its substrate plasminogen, which also binds fibrin. PAI-1
released from platelets is also localized within the thrombus, where
it is associated with vitronectin (107). In humans most of the
PAI-1 in platelets is latent (21,208,209,210). However, there is
enough active PAI-1 in platelets to ensure that the initial fibrin
matrix is not prematurely lysed by tPA activation of plasminogen
(211). PAI-1 deficiency in humans results in a hyperfibrinolytic
state (212), and complete absence of PAI-1 results in
posttraumatic bleeding without other manifestations
(212,213,214).

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

21 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

FIGURE 19-5. Quantitative analysis of angiogenesis in Matrigel


implants. Wild-type, plasminogen activator inhibitor
(PAI)-1deficient or PAI-1transgenic C57/B6J mice were
injected subcutaneously with Matrigel alone or with Matrigel
containing fibroblast growth factor (FGF)-2 and heparin at a
final concentration of 250 ng per mL of FGF-2 and 0.0025 U per
mL of heparin. Wild-type mice were also injected with
FGF-2/heparin and with increasing concentrations of PAI-1.
After 5 to 7 days, the Matrigel implant was harvested and
solubilized in 0.1% Triton X-100 and then centrifuged to
remove particulates. The concentration of hemoglobin in the
supernatant was then determined directly by measuring the
absorbance at 405 nm and by comparing to a standard curve
for purified hemoglobin. A: Hemoglobin content from implants
containing FGF-2 in wild-type, PAI-1deficient (*, P<, 0.05), or
PAI-1transgenic mice (*, P<, 0.05), B: Hemoglobin content in
implants containing FGF-2 with increasing concentrations of
fully active PAI-1. In both panels, n4 for each condition, and
the SE (standard error) are shown. As a control for
background, the hemoglobin levels from implants without
FGF-2 was subtracted form each sample. (From McMahon GA,
Petitclerc E, Stefansson S, et al. Plasminogen activator
inhibitor-1 regulates tumor growth and angiogenesis. J Biol
Chem 2001;276:3396433968, with permission.)

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

22 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

Vascular Wound Healing and Restenosis


Abnormal expression of the fibrinolytic system may also promote
the development of luminal stenosis resulting from arterial
neointima formation after vascular injury. As with tumor growth
and angiogenesis, studies analyzing the role of PAI-1 in neointima
formation and restenosis are variable and are frequently
contradictory. A role for PAI-1 in vascular restenosis was first
suggested by studies where an electrical injury to the femoral
artery was induced in mice genetically deficient in tPA, uPA, or
PAI-1 (215,216). These studies demonstrated that in this model,
neointima formation was primarily the result of smooth muscle cell
migration and proliferation, which, compared to wild-type animals,
was enhanced in the PAI-1 -/ - , reduced in uPA - /- and unchanged in
tPA -/ - mice. These results were similar to those obtained in in vitro
cell migration assays, showing that PAI-1 can inhibit smooth
muscle cell migration by obscuring the cell adhesion site to
vitronectin (167,217). However, work by other groups using
different vascular injury models has yielded opposite conclusions
about the role of PAI-1 in neointima formation. In studies using
either copper-induced (218,219) oxidative vascular injury
(220,221) or carotid ligation (221), PAI-1 was found to inhibit
neointima formation. Consistent with these reports, PAI-1
overexpression by adenovirus transgene transduction was found to
increase vascular stenosis after balloon injury in a rat (222).
The binding of PAI-1 to vitronectin has also been shown to alter
smooth muscle cell migration in different experimental models
(167). Studies with vitronectin - / - mice (221) have shown that
vitronectin deficiency results in a smaller neointima formation
following vascular injury. Likewise, there is no difference between
the smooth muscle cell proliferation in the vitronectin-null mice
compared to wild-type mice, suggesting that vitronectin could
promote neointima formation by enhancing smooth muscle cell
migration. These results raise the possibility that, because
vitronectin can bind simultaneously to fibrin and PAI-1, it can also
promote the inhibition of fibrinolysis by targeting PAI-1 to the
fibrin surface. Therefore, vitronectin may support neointima
formation both by stabilizing fibrin in a PAI-1-dependent manner

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

23 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

and by enhancing the interaction between the provisional matrix


and the smooth muscle cells through integrins.

Atherosclerosis
Although most patients with generalized arterial atherosclerosis
exhibit normal plasma fibrinolytic profiles, the local fibrinolytic
balance in these patients may be severely disturbed. Intravascular
or mural thrombosis is a frequent histologic feature of
atherosclerotic lesions and appears to play a role in the intimal
thickening and fibrosis characteristic of advanced lesions (223). To
determine whether localized alterations in fibrinolytic activity could
influence this process, PAI-1 mRNA expression was evaluated in
segments of severely diseased and relatively normal human
arteries obtained from patients undergoing reconstructive surgery
for aortic occlusive or aneurysmal disease (224) (see Fig. 19-6).
Compared with normal or mildly affected arteries, PAI-1 mRNA
levels in severely atherosclerotic vessels that were analyzed by
Northern blot were considerably increased. In most cases, the level
of PAI-1 gene expression correlated with the degree of
atherosclerosis. Analysis by in situ hybridization demonstrated an
abundance of PAI-1 mRNA-positive cells within the thickened
intima of atherosclerotic arteries, mainly around the base of the
plaque, and in cells scattered within the necrotic material. In
contrast to these results, PAI-1 mRNA was detected primarily
within the luminal endothelial cells of normal-appearing aortic
tissues (224). These observations have been confirmed and
extended (225) in studies showing that both endothelial cells and
smooth muscle cells of apparently normal arteries were positive for
PAI-1 antigen and mRNA. In advanced atherosclerotic lesions,
increased expression of PAI-1 was seen in the smooth muscle cells
within the fibrous cap of the necrotic core. In addition, large
quantities of PAI-1 were present in the extracellular matrix in close
association with elastic lamina and collagen bundles.
PAI-1 has also been found to increase considerably during the
progression from normal vessels to fatty streaks to the developed
atherosclerotic plaque (226). Staining for PAI-1 is strongly
positive, particularly in the areas adjacent to the plaque, whereas
tPA shows the opposite trend, being lowest in lesions with plaque

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

24 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

(226). Therefore, higher concentrations of PAI-1, together with


lower levels of tPA, are characteristic of advanced atheromatous
lesions and alterations in the balance of the fibrinolytic system
that favor its inhibition appear to be a component of the
atherosclerotic process and its common clinical complications.
The role of PAI-1 in the development of atherosclerotic lesions was
also studied in mice lacking either apolipoprotein E (apoE - /- ) or the
LDR-R -/ - These mice develop atheroscleroticlike lesions when fed a
high fat diet. Studies where apoE - / - or LDR-R -/ - mice were crossed
with either transgenic mice overexpressing PAI-1 or mice deficient
in PAI-1 have indicated that, contrary to observations made in
human autopsy specimens (226), PAI-1 expression levels have no
effect on the progression of the disease in the region of the aortic
arch (227). In contrast, a different study, where apoE -/ - mice and
animals deficient in both apoE and PAI-1 were compared, showed a
statistically significant protection from the development of
atherosclerosis at the carotid artery bifurcation in mice lacking
PAI-1 (228).

Cardiovascular and Cerebrovascular Disease


A role for PAI-1 in thrombotic events has been studied in animals
and humans. PAI-1transgenic mice develop age-dependent
coronary artery thrombosis (229), and mice exhibit accelerated
clot-lysis time (230). There is a statistically significant increase in
the PAI-1 expression in mice as a result of stress (231). The
importance of PAI-1 in human vascular disease is suggested by the
finding that human plasma concentrations of PAI-1 increase by
almost 10-fold at the site of injury when platelets are activated
(232). Likewise, a link between increased PAI-1 levels, decreased
fibrinolytic activity, and cardiovascular disease has been suggested
by several studies (233,234,235), and, as was discussed earlier,
increased PAI-1 levels constitute the link between obesity, insulin
resistance, and the resultant increase in the risk of cardiovascular
disease (135,236,237).
P.373
A number of sequence variations in the promoter region of the
PAI-1 gene have been described. One of them, referred to as the

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

25 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

4G/5G polymorphism, is the guanosine deletion/insertion 675 bp


upstream from start of transcription. Although some in vitro
studies have suggested that the 4G allele is associated with higher
PAI-1 activity compared to the 5G allele (238), no association
between PAI-1 genotype (4G/5G polymorphism) and arterial or
venous thrombosis has been found (239), except in patients with
combined protein S deficiency (240). Therefore, there is no reason
to include the study of PAI-1 4G/5G polymorphism as part of the
diagnostic process in patients with cerebrovascular or
cardiovascular disease. Nevertheless, some studies with human
subjects have suggested that patients with 4G/5G polymorphism
and insulin resistance syndrome or sepsis have an increased risk of
myocardial infarction (240) and vascular complications during
systemic infections (118).

FIGURE 19-6. Plasminogen activator inhibitor-1 (PAI-1) gene


expression in the abdominal aortic aneurysm wall. Tissue
samples were analyzed for the expression of PAI-1 mRNA
(messenger ribonucleic acid) by in situ hybridization. A:
Atherosclerotic aneurysm wall. The PAI-1 transcript is
expressed in cells (arrows) aligned at the base of the necrotic
atheroma core, 200, bright field. B: Atherosclerotic aneurysm
wall. Macrophagelike cells expressing PAI-l mRNA (arrows)

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

26 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

within the inflammatory infiltrate, 400, epiluminescence. C:


Atherosclerotic aneurysm wall. PAI-l mRNA expression in
circumferentially arranged cells (arrows), which depict a
cross-section of ringlike structures that are assumed to be
small caliber capillaries (C) within an inflammatory infiltrate,
1,000 bright field. D: Normal aorta. PAI-1 mRNA signal is
detected in luminal endothelium and in a few subintimal cells,
400 epiluminescence (see Color Fig. 19-6). (From
Schneiderman J, Bordin GM, Engelberg I, et al. Expression of
fibrinolytic genes in atherosclerotic abdominal aortic aneurysm
wall. A possible mechanism for aneurysm expansion. J Clin
Invest 1995;96:639645, with permission.)

The role for PAI-1 in cerebral ischemia is less well characterized.


In animal models, PAI-1 mRNA substantially increases in the
ischemic area early after the onset of cerebral ischemia (241,242),
and inactivation of the PAI-1 gene results in a considerable
increase in the volume of the ischemic area following middle
cerebral artery occlusion (243). However, studies in nonhuman
primates have failed to show any link between PAI-1 levels and
neuronal death following cerebral ischemia (242). Likewise, as with
cardiovascular disease, no link has been demonstrated between
4G/5G polymorphism and increased risk of ischemic stroke (244).

Obesity
PAI-1 synthesis has been reported in murine adipocyte cell lines
(128,245,246,247,248), human adipose tissue explants
(236,245,249), and primary cultures of human adipocytes
(250,251) (see Fig. 19-7). There is a considerable correlation
between the amount of visceral fat and plasma levels of PAI-1 in
humans (236,252,253,254) and mice (255,256). Adipose tissue
PAI-1 mRNA levels are enhanced in obese individuals (128), and a
potential role for PAI-1 in obesity has been suggested by the
finding that genetically obese and diabetic (ob/ob) mice crossed
into a PAI-1-deficient background had considerably reduced body
weight and had improved metabolic profiles compared to lean mice
(237). Likewise, nutritionally induced obesity and insulin

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

27 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

resistance were dramatically attenuated in wild-type mice lacking


PAI-1 (257). The improved adiposity and insulin resistance may be
related to the observation that PAI-1-deficient mice fed a high fat
diet had increased metabolic rates and total energy expenditure
compared to wild-type mice and that peroxisome
proliferator-activated receptor (PPAR) and adiponectin were
maintained (257). These observations suggest that PAI-1 has a
direct role in obesity and insulin resistance.
PAI-1 is dramatically upregulated in obesity, and it is now clear
that accelerated atherosclerosis, increased risk for fatal myocardial
infarction, hypertension, insulin resistance, and type 2 diabetes
also frequently accompany this disorder. The possibility that
adipose tissue itself may directly contribute to
P.374
the elevated expression of PAI-1 in obesity has gained
considerable attention. Initial clues for such a hypothesis came
from the observation that the adipose tissue of mice contains
relatively high levels of PAI-1 mRNA (126). Moreover, clinical
studies demonstrated that weight loss due to surgical treatment,
diet, and so on, considerably reduced plasma PAI-1 levels in obese
humans (258). These findings were noteworthy because, in
obesity, the size and number of adipocytes and, therefore, the
amount of adipose tissue mass typically increase severalfold.
Therefore, in obesity, the PAI-1 biosynthetic capacity of adipose
tissue may approach or even exceed that of other tissues.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

28 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

FIGURE 19-7. Localization of plasminogen activator inhibitor-1


(PAI-1) messenger ribonucleic acid (mRNA) in the adipose
tissue of CB6 mice after lipopolysaccharide (LPS) or tumor
necrosis factor (TNF)- treatment. In situ hybridization of
paraffin sections showing vasculature from epididymal fat pads
of untreated mice (A) or from mice treated with LPS (B) or
with TNF-(C) for 3 hours. e, endothelial cells; a, adventitial
cells; s, cells within smooth muscle layers. In situ hybridization
on sections of epididymal fat padcontaining adipocytes and
microvascular endothelial cells from untreated mice (D) or
from mice treated with LPS (E) or TNF- (F) for 3 hours. Some
positive cells are indicated by arrowheads. Slides were exposed
for 8 weeks at 4C and were stained with hematoxylin and
eosin (see Color Fig. 19-7). Original magnification, 400.
(From Samad F, Yamamoto K, Loskutoff DJ. Distribution and

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

29 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

regulation of plasminogen activator inhibitor-1 in murine


adipose tissue in vivo. Induction by tumor necrosis
factor-alpha and lipopolysaccharide. J Clin Invest
1996;97:3746, with permission.)

These initial observations have been extended considerably by


studies in genetically obese (ob/ob) mice (128,255). Plasma PAI-1
activity is approximately fivefold higher in these mice than in their
lean counterparts, and this elevation increases further with age
(255). Moreover, studies with genetically obese and diabetic mice
(ob/ob) lacking the PAI-1 gene (PAI-1 -/ - ) have demonstrated that
elevated PAI-1 is associated with hyperglycemia, hyperinsulinemia,
and insulin resistance syndrome (237).
A variety of observations implicate specific hormones or cytokines
in the increased expression of PAI-1 by adipose
P.375
tissue in obesity. For example, the observations that TNF-, TNF-,
and insulin are elevated in obesity and induce PAI-1 in the plasma
and adipose tissue of lean mice (135) certainly suggest the
involvement of these mediators in the regulation of PAI-1 in
obesity. The observation that inhibiting TNF- in obese mice
considerably decreases plasma PAI-1 antigen and adipose tissue
PAI-1 mRNA (259) strongly supports this hypothesis. Triglycerides
and free fatty acids also may be involved because they stimulate
PAI-1 gene expression in adipocytes. These studies of PAI-1
suggest that these mediators may promote the increased risk for
cardiovascular disease in obesity or type 2 diabetes.

Renal Disease
Although PAI-1 is essentially undetectable in normal kidneys, PAI-1
mRNA and PAI-1 protein levels have been found to be increased in
several renal diseases (260,261,262,263). A role for PAI-1 has
been described in acute renal disorders such as proliferative
glomerulonephritis, renal vasculitis, thrombotic microangiopathy,
and membranous nephropathy, as well as in chronic progressive
renal disease.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

30 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

PAI-1 in Acute Renal Disease. PAI-1 mRNA has been detected in


human and animal models of proliferative glomerulonephritis,
particularly in those cases associated with fibrin deposition (264).
Moreover, it has been observed that inhibition of plasminogen
activators results in a considerable worsening in the severity of the
glomerular injury. PAI-1 reduces glomerular mesangial turnover by
inhibiting plasminogen activators, thereby decreasing plasmin
generation and plasmin-mediated matrix degradation. Because
fibrin accumulation is an important mediator of acute glomerular
injury, strategies designed to enhance fibrinolysis by blocking
PAI-1, or to prevent fibrin formation, might be therapeutic in
patients with glomerulosclerosis. Although the role of PAI-1 in
renal vasculitis is less clear, PAI-1 protein has been identified
together with fibrin deposits in renal biopsy specimens from
patients with focal necrotizing glomerulonephritis due to systemic
lupus erythematosus (265).
In thrombotic microangiopathy, which is a condition that typically
involves fibrin deposition in glomerular capillaries and
extraglomerular arterioles, PAI-1 plays an active role in the
generation of the fibrin thrombi that are formed in response to
glomerular endothelial cell injury. PAI-1 deposition has been
identified in the kidneys of patients with this disorder (266).
Likewise, elevated plasma PAI-1 levels have been associated with
disease outcome in several studies of patients with hemolytic
uremic syndrome (267). Finally, in membranous nephropathy,
PAI-1 transcripts have been found to be abundant (265), and PAI-1
protein colocalizes with vitronectin within the transmembranous
deposits (268).
PAI-1 in Progressive Chronic Renal Disease. In this disease,
there is a pathologic accumulation of extracellular matrix
(269,270). PAI-1 mediates several effects that may facilitate
matrix accumulation through the impairment of matrix turnover.
Likewise, TGF-, which is a critical mediator of renal fibrosis, is a
powerful inducer of PAI-1 expression (271), and several studies
have demonstrated a link between the
reninangiotensinaldosterone cascade, TGF-, and PAI-1 in the
pathophysiology of glomerulosclerosis (269). Angiotensin II
stimulates PAI-1 production (272), and the renoprotective effects

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

31 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

of pharmacologic inhibition of angiotensin II are mediated, at least


in part, by a reduction in PAI-1 expression (273). Likewise, recent
studies using a rat model of mesangioproliferative nephritis, known
as antiThy-1 nephritis, demonstrated that treatment with a
dominant-negative human mutant PAI-1 that binds to matrix
vitronectin but does not inhibit plasminogen activators enhances
plasmin generation, increases matrix turnover, and decreases
matrix accumulation in experimental glomerulonephritis
(274,275,276).

Lung Disease
PAI-1 has been identified as a major deleterious mediator in many
acute and chronic inflammatory lung disorders. Adult respiratory
distress syndrome (ARDS) (277,278), idiopathic pulmonary fibrosis
(IPF) (279,280), sarcoidosis (281), hyperoxic lung injury (282),
and bronchopulmonary dysplasia (283) are all associated with
prominent intraalveolar fibrin deposition and development of
pulmonary fibrosis. Fibrin turnover is tightly regulated by the
concerted action of proteases and antiproteases, and inhibition of
plasmin-mediated proteolysis could account for fibrin accumulation
in lung alveoli. The fibrinolytic activity in bronchoalveolar lavage
(BAL) supernatant fluids obtained from patients with these
diseases has been found to be suppressed (282,283). Moreover,
PAI-1 levels in BAL fluids from patients with ARDS and IPF are
considerably elevated, and it was shown that this PAI-1
upregulation impairs the fibrinolytic capacity of the fluid
(277,278). Bleomycin administration has been used extensively to
study the pathogenesis of pulmonary fibrosis in a variety of animal
models (284). Bleomycin causes a pneumonitis that progresses to
fibrosis in a dose-dependent manner within 2 weeks, while
simultaneously suppressing the fibrinolytic activity of BAL fluid in a
pattern similar to that seen in human inflammatory lung disease
(284). In mice, it was demonstrated that PAI-1 is upregulated after
bleomycin administration and that the PAI-1 expression localizes to
areas of fibrin-rich fibroproliferative lesions (285).
Transgenic mice that either overexpress PAI-1 or are genetically
deficient in PAI-1 have been used to determine whether a
cause-and-effect relation exists between PAI-1 expression and the

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

32 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

development of pulmonary fibrosis. The results of one study


demonstrated a strong relation between PAI-1 gene expression and
the degree of pulmonary fibrosis that followed bleomycin
administration (286). In another study, it was shown that the
lungs of mice exposed to hyperoxia overproduced PAI-1 and that
PAI-1 upregulation impaired fibrinolytic activity in the alveolar
compartment (282). It was found that mice genetically deficient in
PAI-1 did not develop intraalveolar fibrin deposits in response to
hyperoxia. These findings suggest that alterations in the
fibrinolytic environment during inflammatory injury influence the
subsequent development of pulmonary fibrosis and provide
evidence for the pathologic contribution of PAI-1 in this event.
In conclusion, there is a strong link between PAI-1 and human
disease. Currently, there are several ongoing clinical trials testing
the effects of the modification of human PAI-1 in different
diseases. Results of these studies will provide new and definitive
insight into the role of PAI-1 in the pathologic and physiologic
processes.

References
1. Dano K, Andreasen PA, Grondahl-Hansen J, et al.
Plasminogen activators, tissue degradation, and cancer. Adv
Cancer Res 1985;44:139266.
2. Vassalli J-D, Sappino A-P, Belin D. The plasminogen
activator/plasmin system. J Clin Invest 1991;88:10671072.
3. Kohler HP, Grant PJ. Plasminogen-activator inhibitor type 1
and coronary artery disease. N Engl J Med
2000;342:17921801.
4. Saksela O, Rifkin DB. Cell-associated plasminogen activation:
regulation and physiological functions. Annu Rev Cell Biol
1988;4:93126.
5. Hoylaerts M, Rijken DC, Lijnen HR, et al. Kinetics of the

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

33 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

activation of plasminogen by human tissue plasminogen


activator. J Biol Chem 1982; 257:29122919.
6. Ellis V, Dano K. Plasminogen activation by receptor-bound
urokinase. Semin Thromb Hemost 1991;17:194200.
7. Hajjar KA, Hamel NM, Harpel PC, et al. Binding of tissue
plasminogen activator to cultured human endothelial cells. J
Clin Invest 1987;80:17121719.
8. Plow EF, Felez J, Miles LA. Cellular regulation of fibrinolysis.
Thromb Haemost 1991;66:3236.
9. Lawrence DA, Ginsburg D. Plasminogen activator inhibitors.
In: High KA, Roberts HR, eds. Molecular basis of thrombosis and
hemostasis, New York: Marcel Dekker, 1995:517543.
P.376
10. Loskutoff DJ, van Mourik JA, Erickson LA, et al. Detection of
an unusually stable fibrinolytic inhibitor produced by bovine
endothelial cells. Proc Natl Acad Sci U S A 1983;80:29562960.
11. Astedt B, Lecander I, Ny T. The placental type plasminogen
activator PAI-2. Fibrinolysis 1987;1:203208.
12. Geiger M. Protein C inhibitor/plasminogen activator inhibitor
3. Fibrinolysis 1988;2:183188.
13. Scott RW, Bergman BL, Bajpai A, et al. Protease nexin.
Properties and a modified purification procedure. J Biol Chem
1985;260:70297034.
14. Osterwalder T, Contartese J, Stoeckli ET, et al. Neuroserpin,
an axonally secreted serine protease inhibitor. EMBO J
1996;15:29442953.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

34 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

15. Hastings GA, Coleman TA, Haudenschild CC, et al.


Neuroserpin, a brain-associated inhibitor of tissue plasminogen
activator is localized primarily in neurons. Implications for the
regulation of motor learning and neuronal survival. J Biol Chem
1997;272:3306233067.
16. Yepes M, Lawrence DA. Neuroserpin: a selective inhibitor of
tissue-type plasminogen activator in the central nervous
system. Thromb Haemost 2004;91:457464.
17. Emeis JJ, van Hinsbergh VWM, Verheijen JH, et al.
Inhibition of tissue-type plasminogen activator by conditioned
medium from cultured human and porcine vascular endothelial
cells. Biochem Biophys Res Commun 1983;110:392398.
18. Philips M, Juul AG, Thorsen S. Human endothelial cells
produce a plasminogen activator inhibitor and a tissue-type
plasminogen activator-inhibitor complex. Biochim Biophys Acta
1984;802:99110.
19. van Mourik JA, Lawrence DA, Loskutoff DJ. Purification of an
inhibitor plasminogen activator (antiactivator) synthesized by
endothelial cells. J Biol Chem 1984;259:1491414921.
20. Kruithof EKO, Tran-Thang C, Ransijn A, et al.
Demonstration of a fast-acting inhibitor of plasminogen
activators in human plasma. Blood 1984; 64:907913.
21. Erickson LA, Ginsberg MH, Loskutoff DJ. Detection and
partial characterization of an inhibitor of plasminogen activator
in human platelets. J Clin Invest 1984;74:14651472.
22. Erickson LA, Hekman CM, Loskutoff DJ. Denaturant-induced
stimulation of the beta-migrating plasminogen activator
inhibitor in endothelial cells and serum. Blood
1986;68:12981305.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

35 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

23. Ny T, Sawdey M, Lawrence D, et al. Cloning and sequence


of a cDNA coding for the human beta-migrating
endothelial-cell-type plasminogen activator inhibitor. Proc Natl
Acad Sci U S A 1986;83:67766780.
24. Chmielewska J, Rnby M, Wiman B. Evidence for a rapid
inhibitor to tissue plasminogen activator in plasma. Thromb Res
1983;31:427436.
25. Kruithof EKO, Ransijn A, Bachmann F. Inhibition of tissue
plasminogen activator by human plasma. In: Davidson JF,
Bachmann F, Bouvier CA, et al, eds. Progress in Fibrinolysis
Volume VI, Edinburgh: Churchill Livingstone, 1983:365369.
26. Gettins PGW, Patston PA, Olson ST. Serpins: structure,
function and biology. Austin, Texas: R.G. Landes Company,
1996.
27. Doolittle RF. Angiotensinogen is related to the
antitrypsin-antithrombin-ovalbumin family. Science
1983;222:417419.
28. Irving JA, Pike RN, Lesk AM, et al. Phylogeny of the serpin
superfamily: implications of patterns of amino acid conservation
for structure and function. Genome Res 2000;10:18451864.
29. Irving JA, Steenbakkers PJ, Lesk AM, et al. Serpins in
prokaryotes. Mol Biol Evol 2002;19:18811890.
30. Cohen AB, Gruenke LD, Craig JC, et al. Specific lysine
labeling by

18

during alkaline cleavage of the

-antitrypsin-trypsin complex. Proc Natl Acad Sci U S A


1977;74:43114314.
31. Wiman B, Collen D. On the mechanism of the reaction

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

36 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

between human alpha 2-antiplasmin and plasmin. J Biol Chem


1979;254:92919297.
32. Levin EG. Latent tissue plasminogen activator produced by
human endothelial cells in culture: evidence for an
enzyme-inhibitor complex. Proc Natl Acad Sci U S A
1983;80:68046808.
33. Lawrence DA, Ginsburg D, Day DE, et al. Serpin-protease
complexes are trapped as stable acyl-enzyme intermediates. J
Biol Chem 1995;270: 2530925312.
34. Huber R, Carrell RW. Implications of the three-dimensional
structure of alpha 1-antitrypsin for structure and function of
serpins. Biochem 1989; 28:89518966.
35. Shubeita HE, Cottey TL, Franke AE, et al. Mutational and
immunochemical analysis of plasminogen activator inhibitor 1. J
Biol Chem 1990;265: 1837918385.
36. York JD, Li P, Gardell SJ. Combinatorial mutagenesis of the
reactive site region in plasminogen activator inhibitor I. J Biol
Chem 1991;266:84958500.
37. Sherman PM, Lawrence DA, Yang AY, et al. Saturation
mutagenesis of the plasminogen activator inhibitor-1 reactive
center. J Biol Chem 1992;267: 75887595.
38. Schechter I, Berger A. On the size of the active site in
proteases. I. Papain. Biochem Biophys Res Commun
1967;27:157162.
39. Wilczynska M, Fa M, Ohlsson PI, et al. The inhibition
mechanism of serpins: evidence that the mobile reactive center
loop is cleaved in the native protease-inhibitor complex. J Biol
Chem 1995;270:2965229655.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

37 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

40. Lawrence DA, Olson ST, Muhammad S, et al. Partitioning of


serpin-proteinase reactions between stable inhibition and
substrate cleavage is regulated by the rate of serpin reactive
center loop insertion into beta-sheet A. J Biol Chem
2000;275:58395844.
41. Hagglof P, Bergstrom F, Wilczynska M, et al. The
reactive-center loop of active PAI-1 is folded close to the
protein core and can be partially inserted. J Mol Biol
2004;335:823832.
42. Lawrence DA, Strandberg L, Ericson J, et al.
Structure-function studies of the SERPIN plasminogen activator
inhibitor type 1: analysis of chimeric strained loop mutants. J
Biol Chem 1990;265:2029320301.
43. Huntington JA, Read RJ, Carrell RW. Structure of a
serpin-protease complex shows inhibition by deformation.
Nature 2000;407:923926.
44. Huntington JA, Carrell RW. The serpins: nature's molecular
mousetraps. Sci Prog 2001;84:125136.
45. Andreasen PA, Riccio A, Welinder KG, et al. Plasminogen
activator inhibitor type-1: reactive center and amino-terminal
heterogeneity determined by protein and cDNA sequencing.
FEBS Lett 1986;209:213218.
46. Pannekoek H, Veerman H, Lambers H, et al. Endothelial
plasminogen activator inhibitor (PAI): a new member of the
Serpin gene family. EMBO J 1986;5:25392544.
47. Ginsburg D, Zeheb R, Yang AY, et al. cDNA cloning of
human plasminogen activator-inhibitor from endothelial cells. J
Clin Invest 1986;78:16731680.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

38 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

48. Lawrence D, Strandberg L, Grundstrm T, et al. Purification


of active human plasminogen activator inhibitor 1 from
Escherichia coli. Comparison with natural and recombinant
forms purified from eucaryotic cells. Eur J Biochem
1989;186:523533.
49. Franke AE, Danley DE, Kaczmarek FS, et al. Expression of
human plasminogen activator inhibitor type-1 (PAI-1) in
Escherichia coli as a soluble protein comprised of active and
latent forms. Isolation and crystallization of latent PAI-1.
Biochim Biophys Acta 1990;1037:1623.
50. Reilly TM, Seetharam R, Duke JL, et al. Purification and
characterization of recombinant plasminogen activator
inhibitor-1 from Escherichia coli. J Biol Chem
1990;265:95709574.
51. Keijer J, Ehrlich HJ, Linders M, et al. Vitronectin governs
the interaction between plasminogen activator inhibitor 1 and
tissue-type plasminogen activator. J Biol Chem
1991;266:1070010707.
52. Seetharam R, Dwivedi AM, Duke JL, et al. Purification and
characterization of active and latent forms of recombinant
plasminogen activator inhibitor 1 produced in Escherichia coli.
Biochem 1992;31:98779882.
53. Sherman PM, Lawrence DA, Verhamme IM, et al.
Identification of tPA-specific plasminogen activator inhibitor-1
mutants: evidence that second sites of interaction contribute to
target specificity. J Biol Chem 1995;270: 93019306.
54. Stefansson S, Yepes M, Gorlatova N, et al. Mutants of
plasminogen activator inhibitor-1 designed to inhibit neutrophil
elastase and cathepsin G are more effective in vivo than their
endogenous inhibitors. J Biol Chem 2004;279:29981-29987.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

39 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

55. Kvassman JO, Verhamme I, Shore JD. Inhibitory mechanism


of serpins: loop insertion forces acylation of plasminogen
activator by plasminogen activator inhibitor-1. Biochem
1998;37:1549115502.
56. Olson ST, Swanson R, Day D, et al. Resolution of Michaelis
complex, acylation, and conformational change steps in the
reactions of the serpin, plasminogen activator inhibitor- 1, with
tissue plasminogen activator and trypsin. Biochem
2001;40:1174211756.
57. Ibarra CA, Blouse GE, Christian TD, et al. The contribution
of the exosite residues of plasminogen activator inhibitor-1 to
proteinase inhibition. J Biol Chem 2004;279:36433650.
58. Hekman CM, Loskutoff DJ. Endothelial cells produce a latent
inhibitor of plasminogen activators that can be activated by
denaturants. J Biol Chem 1985;260:1158111587.
59. Levin EG, Santell L. Conversion of the active to latent
plasminogen activator inhibitor from human endothelial cells.
Blood 1987;70:10901098.
60. Mottonen J, Strand A, Symersky J, et al. Structural basis of
latency in plasminogen activator inhibitor-1. Nature
1992;355:270273.
61. Sharp AM, Stein PE, Pannu NS, et al. The active
conformation of plasminogen activator inhibitor 1, a target for
drugs to control fibrinolysis and cell adhesion. Structure Fold
Des 1999;7:111118.
62. Lawrence DA, Olson ST, Palaniappan S, et al. Engineering
plasminogen activator inhibitor-1 (PAI-1) mutants with
increased functional stability. Biochem 1994;33:36433648.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

40 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

63. Lawrence DA, Olson ST, Palaniappan S, et al. Serpin


reactive-center loop mobility is required for inhibitor function
but not for enzyme recognition. J Biol Chem
1994;269:2765727662.
64. Hekman CM, Loskutoff DJ. Bovine plasminogen activator
inhibitor 1: specificity determinations and comparison of the
active, latent, and guanidine-activated forms. Biochem
1988;27:29112918.
65. Lindahl TL, Sigurdardttir O, Wiman B. Stability of
plasminogen activator inhibitor 1 (PAI-1). Thromb Haemost
1989;62:748751.
66. Lambers JW, Cammenga M, Konig BW, et al. Activation of
human endothelial cell-type plasminogen activator inhibitor
(PAI-1) by negatively charged phospholipids. J Biol Chem
1987;262:1749217496.
67. Wun T-C, Palmier MO, Siegel NR, et al. Affinity purification
of active plasminogen activator inhibitor- 1 (PAI-1) using
immobilized anhydrourokinase. J Biol Chem
1989;264:78627868.
68. Lawrence DA, Loskutoff DJ. Inactivation of plasminogen
activator inhibitor by oxidants. Biochem 1986;25:63516355.
69. Strandberg L, Lawrence DA, Johansson LB-A, et al. The
oxidative inactivation of plasminogen activator inhibitor type 1
results from a conformational change in the molecule and does
not require the involvement of the methionine. J Biol Chem
1991;266:1385213858.
P.377
70. Weiss SJ, Regiani S. Neutrophils degrade subendothelial
matrices in the presence of alpha-1-proteinase inhibitor.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

41 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

Cooperative use of lysosomal proteinases and oxygen


metabolites. J Clin Invest 1984;73: 12971303.
71. Declerck PJ, De Mol M, Vaughan DE, et al. Identification of
a conformationally distinct form of plasminogen activator
inhibitor-1, acting as a non-inhibitory substrate for tissue-type
plasminogen activator. J Biol Chem 1992;267:1169311696.
72. Urano T, Strandberg L, Johansson LB, et al. A substrate-like
form of plasminogen-activator-inhibitor type 1conversions
between different forms by sodium dodecyl sulphate. Eur J
Biochem 1992;209:985992.
73. Wu K, Urano T, Ihara H, et al. The cleavage and inactivation
of plasminogen activator inhibitor type 1 by neutrophil elastase:
the evaluation of its physiologic relevance in fibrinolysis. Blood
1995;86:10561061.
74. Lawrence DA. The serpin-proteinase complex revealed. Nat
Struct Biol 1997;4:339341.
75. Wilczynska M, Fa M, Karolin J, et al. New structural insights
into native serpin-protease complexes reveal the inhibitory
mechanism of serpins. Nat Struct Biol 1997;4:354357.
76. Sprengers ED, Kluft C. Plasminogen activator inhibitors.
Blood 1987;69: 381387.
77. Stefansson S, Petitclerc E, Wong MK, et al. Inhibition of
Angiogenesis in vivo by Plasminogen activator inhibitor-1. J Biol
Chem 2001;276:81358141.
78. Booth NA. Fibrinolysis and thrombosis. Baillieres Best Pract
Res Clin Haematol 1999;12:423433.
79. Huber K. Plasminogen activator inhibitor type-1 (part one):

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

42 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

basic mechanisms, regulation, and role for thromboembolic


disease. J Thromb Thrombolysis 2001;11:183193.
80. Stefansson S, McMahon GA, Petitclerc E, et al. Plasminogen
activator inhibitor-1 in tumor growth, angiogenesis and vascular
remodeling. Curr Pharma Design 2003;9:15451564.
81. Lawrence DA, Berkenpas MB, Palaniappan S, et al.
Localization of vitronectin binding domain in plasminogen
activator inhibitor-1. J Biol Chem 1994;269:1522315228.
82. Xu Z, Balsara RD, Gorlatova NV, et al. Conservation of
critical functional domains in murine plasminogen activator
inhibitor-1. J Biol Chem 2004; 279:1791417920.
83. Jensen JK, Wind T, Andreasen PA. The vitronectin binding
area of plasminogen activator inhibitor-1, mapped by
mutagenesis and protection against an inactivating
organochemical ligand. FEBS Lett 2002;521:9194.
84. Zhou A, Huntington JA, Pannu NS, et al. How vitronectin
binds PAI-1 to modulate fibrinolysis and cell migration. Nat
Struct Biol 2003;10:541544.
85. Stefansson S, Muhammad S, Cheng XF, et al. Plasminogen
activator inhibitor-1 contains a cryptic high affinity binding site
for the low density lipoprotein receptor-related protein. J Biol
Chem 1998;273:63586366.
86. Horn IR, van den Berg BM, Moestrup SK, et al. Plasminogen
activator inhibitor 1 contains a cryptic high affinity receptor
binding site that is exposed upon complex formation with
tissue-type plasminogen activator. Thromb Haemost
1998;80:822828.
87. Tomasini BR, Mosher DF. Vitronectin. Prog Hemost Thromb

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

43 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

1991;10: 269305.
88. Seiffert D. Constitutive and regulated expression of
vitronectin. Histol Histopathol 1997;12:787797.
89. Podor TJ, Campbell S, Chindemi P, et al. Incorporation of
vitronectin into fibrin clots. Evidence for a binding interaction
between vitronectin and gamma A/gamma' fibrinogen. J Biol
Chem 2002;277:75207528.
90. Seiffert D, Smith JW. The cell adhesion domain in plasma
vitronectin is cryptic. J Biol Chem 1997;272:1370513710.
91. Keijer J, Linders M, Wegman JJ, et al. On the target
specificity of plasminogen activator inhibitor 1: the role of
heparin, vitronectin, and the reactive site. Blood
1991;78:12541261.
92. Naski MC, Lawrence DA, Mosher DF, et al. Kinetics of
inactivation of -thrombin by plasminogen activator inhibitor-1:
comparison of the effects of native and urea-treated forms of
vitronectin. J Biol Chem 1993; 268:1236712372.
93. Lawrence DA, Palaniappan S, Stefansson S, et al.
Characterization of the binding of different conformational
forms of plasminogen activator inhibitor-1 to vitronectin:
implications for the regulation of pericellular proteolysis. J Biol
Chem 1997;272:76767680.
94. Grondahl-Hansen J, Lund LR, Ralfkir E, et al.
Urokinase-tissue-type plasminogen activators in keratinocytes
during wound reepithelialization. J Invest Dermatol
1988;90:790795.
95. Yebra M, Parry GCN, Stromblad S, et al. Requirement of
receptor-bound urokinase-type plasminogen activator for

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

44 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

integrin v 5 -directed cell migration. J Biol Chem


1996;271:2939329399.
96. Carriero MV, Del Vecchio S, Capozzoli M, et al. Urokinase
receptor interacts with alpha(v)beta5 vitronectin receptor,
promoting urokinase-dependent cell migration in breast cancer.
Cancer Res 1999;59:53075314.
97. Tarui T, Mazar AP, Cines DB, et al. Urokinase-type
plasminogen activator receptor (cd87) is a ligand for integrins
and mediates cell-cell interaction. J Biol Chem
2001;276:39833990.
98. Chapman HA, Wei Y. Protease crosstalk with integrins: the
urokinase receptor paradigm. Thromb Haemost
2001;86:124129.
99. Lindahl T, Wiman B. Purification of high and low molecular
weight plasminogen activator inhibitor 1 from fibrosarcoma
cell-line HT 1080 conditioned medium. Biochim Biophys Acta
1989;994:253257.
100. Ehrlich HJ, Keijer J, Preissner KT, et al. Functional
interaction of plasminogen activator inhibitor type 1 (PAI-1)
and heparin. Biochem 1991;30: 10211028.
101. Ehrlich HJ, Gebbink RK, Keijer J, et al. Elucidation of
structural requirements on plasminogen activator inhibitor 1 for
binding to heparin. J Biol Chem 1992;267:1160611611.
102. Braaten JV, Handt S, Jerome WG, et al. Regulation of
fibrinolysis by platelet-released plasminogen activator inhibitor
1: light scattering and ultrastructural examination of lysis of a
model platelet-fibrin thrombus. Blood 1993;81:12901299.
103. Wagner OF, de Vries C, Hohmann C, et al. Interaction

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

45 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

between plasminogen activator inhibitor type 1 (PAI- 1) bound


to fibrin and either tissue-type plasminogen activator (t-PA) or
urokinase-type plasminogen activator (u-PA). J Clin Invest
1989;84:647655.
104. Keijer J, Linders M, van Zonneveld A-J, et al. The
interaction of plasminogen activator inhibitor 1 with
plasminogen activators (tissue-type and urokinase-type) and
fibrin: localization of interaction sites and physiologic
relevance. Blood 1991;78:401409.
105. Reilly CF, Hutzelmann JE. Plasminogen activator
inhibitor-1 binds to fibrin and inhibits tissue-type plasminogen
activator-mediated fibrin dissolution. J Biol Chem
1992;267:1712817135.
106. Podor TJ, Shaughnessy SG, Blackburn MN, et al. New
insights into the size and stoichiometry of the plasminogen
activator inhibitor type-1. vitronectin complex. J Biol Chem
2000;275:2540225410.
107. Podor TJ, Peterson CB, Lawrence DA, et al. Type 1
plasminogen activator inhibitor binds to fibrin via vitronectin. J
Biol Chem 2000;275: 1978819794.
108. Klinger KW, Winqvist R, Riccio A, et al. Plasminogen
activator inhibitor type 1 gene is located at region q21.3-q22 of
chromosome 7 and genetically linked with cystic fibrosis. Proc
Natl Acad Sci U S A 1987;84: 85488552.
109. Loskutoff DJ, Linders M, Keijer J, et al. Structure of the
human plasminogen activator inhibitor 1 gene: nonrandom
distribution of introns. Biochem 1987;26:37633768.
110. Strandberg L, Lawrence D, Ny T. The organization of the
human plasminogen-activator-inhibitor 1 gene. Eur J Biochem

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

46 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

1988;176:609616.
111. Follo M, Ginsburg D. Structure and expression of the
human gene encoding plasminogen activator inhibitor, PAI-1.
Gene 1989;84:447453.
112. Bosma PJ, van den Berg EA, Kooistra T, et al. Human
plasminogen activator inhibitor-1 gene: promoter and structural
gene nucleotide sequences. J Biol Chem 1988;263:91299141.
113. McGrogan M, Kennedy J, Golini F. Structure of the human
protease nexin gene and expression of recombinant forms of
PN-1. In: Festoff BW, ed. Serine proteases and their serpin
inhibitors in the nervous system. New York: Plenum Press,
1990:147161.
114. Bosma PJ, Kooistra T, Siemieniak DR, et al. Further
characterization of the 5-flanking DNA of the gene encoding
human plasminogen activator inhibitor-1. Gene
1991;100:261266.
115. Berger P, Kozlov SV, Krueger SR, et al. Structure of the
mouse gene for the serine protease inhibitor neuroserpin
(PI12). Gene 1998;214:2533.
116. van Zonneveld A-J, Curriden SA, Loskutoff DJ. Type 1
plasminogen activator inhibitor gene: functional analysis and
glucocorticoid regulation of its promoter. Proc Natl Acad Sci U S
A 1988;85:55255529.
117. Riccio A, Lund LR, Sartorio R, et al. The regulatory region
of the human plasminogen activator inhibitor type-1 (PAI-1)
gene. Nucleic Acids Res 1988;16:28052824.
118. Dawson SJ, Wiman B, Hamsten A, et al. The two allele
sequences of a common polymorphism in the promoter of the

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

47 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

plasminogen activator inhibitor-1 (PAI-1) gene respond


differently to interleukin-1 in HepG2 cells. J Biol Chem
1993;268:1073910745.
119. Dawson SJ, Hamsten A, Wiman B, et al. Genetic variation
at the plasminogen activator inhibitor-1 locus is associated with
altered levels of plasma plasminogen activator inhibitor-1
activity. Arterioscler Thromb 1991;11:183190.
120. van den Berg EA, Sprengers ED, Jaye M, et al. Regulation
of plasminogen activator inhibitor-1 mRNA in human endothelial
cells. Thromb Haemost 1988;60:6367.
121. Konkle BA, Ginsburg D. The addition of endothelial cell
growth factor and heparin to human umbilical vein endothelial
cell cultures decreases plasminogen activator inhibitor-1
expression. J Clin Invest 1988;82:579585.
122. Irigoyen JP, Munoz-Canoves P, Montero L, et al. The
plasminogen activator system: biology and regulation. Cell Mol
Life Sci 1999;56:104132.
123. Heaton JH, Dlakic WM, Gelehrter TD. Posttranscriptional
regulation of PAI-1 gene expression. Thromb Haemost
2003;89:959966.
124. Kruithof EKO, Gudinchet A, Bachmann F. Plasminogen
activator inhibitor 1 and plasminogen activator inhibitor 2 in
various disease states. Thromb Haemost 1988;59:712.
125. Colucci M, Paramo JA, Collen D. Generation in plasma of a
fast-acting inhibitor of plasminogen activator in response to
endotoxin stimulation. J Clin Invest 1985;75:818824.
126. Sawdey MS, Loskutoff DJ. Regulation of murine type 1
plasminogen activator inhibitor gene expression in vivo. Tissue

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

48 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

specificity and induction by lipopolysaccharide, tumor necrosis


factor-, and transforming growth factor-. J Clin Invest
1991;88:13461353.
P.378
127. Fearns C, Loskutoff DJ. Induction of plasminogen activator
inhibitor 1 gene expression in murine liver by
lipopolysaccharide. Cellular localization and role of endogenous
tumor necrosis factor-alpha. Am J Pathol 1997;150: 579590.
128. Samad F, Yamamoto K, Loskutoff DJ. Distribution and
regulation of plasminogen activator inhibitor-1 in murine
adipose tissue in vivo. Induction by tumor necrosis factor-alpha
and lipopolysaccharide. J Clin Invest 1996;97:3746.
129. Sawdey M, Podor TJ, Loskutoff DJ. Regulation of type 1
plasminogen activator inhibitor gene expression in cultured
bovine aortic endothelial cells. J Biol Chem
1989;264:1039610401.
130. Loskutoff DJ. Regulation of PAI-1 gene expression.
Fibrinolysis 1991; 5:197206.
131. Fattal PG, Schneider DJ, Sobel BE, et al.
Post-transcriptional regulation of expression of plasminogen
activator inhibitor type 1 on mRNA by insulin and insulin-like
growth factor 1. J Biol Chem 1992;267:1241212415.
132. Konkle BA, Schuster SJ, Kelly MD, et al. Plasminogen
activator inhibitor-1 messenger RNA expression is induced in
rat hepatocytes in vivo by dexamethasone. Blood
1992;79:26362642.
133. Loskutoff DJ, Sawdey M, Keeton M, et al. Regulation of
PAI-1 gene expression in vivo. Thromb Haemost
1993;70:135137.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

49 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

134. Simpson AJ, Booth NA, Moore NR, et al. Distribution of


plasminogen activator inhibitor (PAI-1) in tissues. J Clin Pathol
1991;44:139143.
135. Loskutoff DJ, Samad F. The adipocyte and hemostatic
balance in obesity: studies of PAI-1. Arterioscler Thromb Vasc
Biol 1998;18:16.
136. Lawrence DA, Ginsburg D. Gene expression and function of
plasminogen activator inhibitor-1. In: Glas-Greenwalt P, ed.
Fibrinolysis in disease: molecular and Hhemovascular aspects of
fibrinolysis, Boca Raton: CRC Press, 1995:2129.
137. Mutch NJ, Wilson HM, Booth NA. Plasminogen activator
inhibitor-1 and haemostasis in obesity. Proc Nutr Soc
2001;60:341347.
138. Unkeless J, Dano K, Kellerman GM, et al. Fibrinolysis
associated with oncogenic transformation. Partial purification
and characterization of the cell factor, a plasminogen activator.
J Biol Chem 1974;249:42954305.
139. Astrup T. Cell-induced fibrinolysis: a fundamental process.
In: Reich E, Rifkin DB, Shaw E, eds. Proteases and Biological
Control, Cold Spring Harbor: Cold Spring Harbor Laboratory,
1975:343355.
140. Duffy MJ. Plasminogen activators and cancer. Blood Coagul
Fibrinolysis 1990;1:681687.
141. Schmitt M, Wilhelm O, Janicke F, et al. Urokinase-type
plasminogen activator (uPA) and its receptor (CD87): a new
target in tumor invasion and metastasis. J Obstet Gynaecol
1995;21:151165.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

50 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

142. Hofmann R, Lehmer A, Buresch M, et al. Clinical relevance


of urokinase plasminogen activator, its receptor, and its
inhibitor in patients with renal cell carcinoma. Cancer
1996;78:487492.
143. Quax PH, van Muijen GN, Weening-Verhoeff EJ, et al.
Metastatic behavior of human melanoma cell lines in nude mice
correlates with urokinase-type plasminogen activator, its type-1
inhibitor, and urokinase-mediated matrix degradation. J Cell
Biol 1991;115:191199.
144. Sier CF, Vloedgraven HJ, Ganesh S, et al. Inactive
urokinase and increased levels of its inhibitor type 1 in
colorectal cancer liver metastasis. Gastroenterology
1994;107:14491456.
145. Pedersen H, Brunner N, Francis D, et al. Prognostic impact
of urokinase, urokinase receptor, and type 1 plasminogen
activator inhibitor in squamous and large cell lung cancer
tissue. Cancer Res 1994;54:46714675.
146. Nekarda H, Siewert JR, Schmitt M, et al.
Tumour-associated proteolytic factors uPA and PAI-1 and
survival in totally resected gastric cancer. Lancet
1994;343:117.
147. Pyke C, Kristensen P, Ralfkiaer E, et al. The plasminogen
activation system in human colon cancer: messenger RNA for
the inhibitor PAI-1 is located in endothelial cells in the tumor
stroma. Cancer Res 1991;51: 40674071.
148. Pyke C, Kristensen P, Ralfkiaer E, et al. Urokinase-type
plasminogen activator is expressed in stromal cells and its
receptor in cancer cells at invasive foci in human colon
adenocarcinomas. Am J Pathol 1991;138:10591067.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

51 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

149. Kristensen P, Pyke C, Lund LR, et al. Plasminogen


activator inhibitor-type 1 in Lewis lung carcinoma.
Histochemistry 1990;93:559566.
150. Rabbani SA, Xing RH. Role of urokinase (uPA) and its
receptor (uPAR) in invasion and metastasis of
hormone-dependent malignancies. Int J Oncol
1998;12:911920.
151. Konno H, Baba M, Shoji T, et al. Cyclooxygenase-2
expression correlates with uPAR levels and is responsible for
poor prognosis of colorectal cancer. Clin Exp Metastasis
2002;19:527534.
152. Pepper MS, Montesano R. Proteolytic balance and capillary
morphogenesis. Cell Differ Dev 1990;32:319327.
153. Romer J, Lund LR, Eriksen J, et al. Differential expression
of urokinase-type plasminogen activator and its type-1 inhibitor
during healing of mouse skin wounds. J Invest Dermatol
1991;97:803811.
154. Lang IM, Moser KM, Schleef RR. Elevated expression of
urokinase-like plasminogen activator and plasminogen activator
inhibitor type 1 during the vascular remodeling associated with
pulmonary thromboembolism. Arterioscler Thromb Vasc Biol
1998;18:808815.
155. Erickson CA, Isseroff RR. Plasminogen activator activity is
associated with neural crest cell motility in tissue culture. J Exp
Zool 1989;251: 123133.
156. Kirchheimer JC, Remold HG. Endogenous receptor-bound
urokinase mediates tissue invasion of human monocytes. J
Immunol 1989;143: 26342639.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

52 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

157. Wilson EL, Jacobs P, Dowdle EB. The secretion of


plasminogen activators by human myeloid leukemic cells in
vitro. Blood 1983;61:568574.
158. Levin EG, Fair DS, Loskutoff DJ. Human hepatoma cell line
plasminogen activator. J Lab Clin Med 1983;102:500508.
159. Hsu DW, Efird JT, Hedley-Whyte ET. Prognostic role of
urokinase-type plasminogen activator in human gliomas. Am J
Pathol 1995;147:114123.
160. Ossowski L. Invasion of connective tissue by human
carcinoma cell lines: requirement for urokinase, urokinase
receptor, and interstitial collagenase. Cancer Res
1992;52:67546760.
161. Coen D, Bottazzi B, Bini A, et al. Plasminogen activator
activity of metastatic variants from a murine fibrosarcoma:
effect of thrombin in vitro. Int J Cancer 1983;32:6770.
162. Yu HR, Schultz RM. Relationship between secreted
urokinase plasminogen activator activity and metastatic
potential in murine B16 cells transfected with human urokinase
sense and antisense genes. Cancer Res 1990;50:76237633.
163. Hearing VJ, Law LW, Corti A, et al. Modulation of
metastatic potential by cell surface urokinase of murine
melanoma cells. Cancer Res 1988;48: 12701278.
164. Ossowski L. Plasminogen activator dependent pathways in
the dissemination of human tumor cells in the chick embryo.
Cell 1988;52:321328.
165. Mignatti P, Robbins E, Rifkin DB. Tumor invasion through
the human amniotic membrane: requirement for a proteinase

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

53 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

cascade. Cell 1986;47: 487498.


166. Stefansson S, Lawrence DA. Old dogs and new tricks,
proteases, inhibitors, and cell migration. Sci STKE
2003;189;pe24.
167. Stefansson S, Lawrence DA. The serpin PAI-1 inhibits cell
migration by blocking integrin v 3 binding to vitronectin.
Nature 1996;383:441443.
168. Kjoller L, Kanse SM, Kirkegaard T, et al. Plasminogen
activator inhibitor-1 represses integrin- and
vitronectin-mediated cell migration independently of its function
as an inhibitor of plasminogen activation. Exp Cell Res
1997;232:420429.
169. Giannelli G, Falk-Marzillier J, Schiraldi O, et al. Induction
of cell migration by matrix metalloprotease-2 cleavage of
laminin-5. Science 1997;277: 225228.
170. Xu J, Rodriguez D, Petitclerc E, et al. Proteolytic exposure
of a cryptic site within collagen type IV is required for
angiogenesis and tumor growth in vivo. J Cell Biol
2001;154:10691079.
171. Hangai M, Kitaya N, Xu J, et al. Matrix
metalloproteinase-9-dependent exposure of a cryptic migratory
control site in collagen is required before retinal angiogenesis.
Am J Pathol 2002;161:14291437.
172. Sato Y, Tsuboi R, Lyons R, et al. Characterization of the
activation of latent TGF-beta by co-cultures of endothelial cells
and pericytes or smooth muscle cells: a self-regulating system.
J Cell Biol 1990;111:757763.
173. Lyons RM, Gentry LE, Purchio AF, et al. Mechanism of

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

54 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

activation of latent recombinant transforming growth factor


beta 1 by plasmin. J Cell Biol 1990;110:13611367.
174. Saksela O, Rifkin DB. Release of basic fibroblast growth
factor-heparan sulfate complexes from endothelial cells by
plasminogen activator-mediated proteolytic activity. J Cell Biol
1990;110:767775.
175. Ribatti D, Leali D, Vacca A, et al. In vivo angiogenic
activity of urokinase: role of endogenous fibroblast growth
factor-2. J Cell Sci 1999;112(Pt 23):42134221.
176. Houck KA, Leung DW, Rowland AM, et al. Dual regulation
of vascular endothelial growth factor bioavailability by genetic
and proteolytic mechanisms. J Biol Chem
1992;267:2603126037.
177. Matsuno H, Kozawa O, Yoshimi N, et al. Lack of
alpha2-antiplasmin promotes pulmonary heart failure via
overrelease of VEGF after acute myocardial infarction. Blood
2002;100:24872493.
178. Pllnen J, Hedman K, Nielsen LS, et al. Ultrastructural
localization of plasma membrane-associated urokinase-type
plasminogen activator at focal contacts. J Cell Biol
1988;106:8795.
179. Hebert CA, Baker JB. Linkage of extracellular plasminogen
activator to the fibroblast cytoskeleton: colocalization of cell
surface urokinase with vinculin. J Cell Biol
1988;106:12411247.
180. Bastholm L, Nielsen MH, De Mey J, et al. Confocal
fluorescence microscopy of urokinase plasminogen activator
receptor and cathepsin D in human MDA-MB-231 breast cancer
cells migrating in reconstituted basement membrane. Biotech

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

55 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

Histochem 1994;69:6167.
181. Wei Y, Waltz DA, Rao N, et al. Identification of the
urokinase receptor as an adhesion receptor for vitronectin. J
Biol Chem 1994;269:3238032388.
182. Wei Y, Eble JA, Wang Z, et al. Urokinase receptors
promote beta1 integrin function through interactions with
integrin alpha3beta1. Mol Biol Cell 2001;12:29752986.
183. Simon DI, Rao NK, Xu H, et al. Mac-1 (CD11b/CD18) and
the urokinase receptor (CD87) form a functional unit on
monocytic cells. Blood 1996;88:31853194.
184. Okumura Y, Kamikubo Y, Curriden SA, et al. Kinetic
analysis of the interaction between vitronectin and the
urokinase receptor. J Biol Chem 2002;277:93959404.
185. Deng G, Curriden SA, Wang S, et al. Is plasminogen
activator inhibitor-1 the molecular switch that governs
urokinase receptor-mediated cell adhesion and release? J Cell
Biol 1996;134:15631571.
186. Kanse SM, Kost C, Wilhelm OG, et al. The urokinase
receptor is a major vitronectin-binding protein on endothelial
cells. Exp Cell Res 1996;224: 344353.
P.379
187. Waltz DA, Natkin LR, Fujita RM, et al. Plasmin and
plasminogen activator inhibitor type 1 promote cellular motility
by regulating the interaction between the urokinase receptor
and vitronectin. J Clin Invest 1997;100: 5867.
188. Czekay RP, Aertgeerts K, Curriden SA, et al. Plasminogen
activator inhibitor-1 detaches cells from extracellular matrices
by inactivating integrins. J Cell Biol 2003;160:781791.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

56 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

189. Inyang AL, Tobelem G. Tissue-plasminogen activator


stimulates endothelial cell migration in wound assays. Biochem
Biophys Res Commun 1990;171:13261332.
190. Kirchheimer JC, Binder BR, Remold HG. Matrix-bound
plasminogen activator inhibitor type 1 inhibits the invasion of
human monocytes into interstitial tissue. J Immunol
1990;145:15181522.
191. Palmieri D, Lee JW, Juliano RL, et al. Plasminogen
activator inhibitor-1 and -3 increase cell adhesion and motility
of MDA-MB-435 breast cancer cells. J Biol Chem
2002;277:4095040957.
192. Ciambrone GJ, McKeown-Longo PJ. Plasminogen activator
inhibitor type I stabilizes vitronectin- dependent adhesions in
HT-1080 cells. J Cell Biol 1990;111:21832195.
193. Webb DJ, Thomas KS, Gonias SL. Plasminogen activator
inhibitor 1 functions as a urokinase response modifier at the
level of cell signaling and thereby promotes MCF-7 cell growth.
J Cell Biol 2001;152:741752.
194. Degryse B, Sier CF, Resnati M, et al. PAI-1 inhibits
urokinase-induced chemotaxis by internalizing the urokinase
receptor. FEBS Lett 2001;505: 249254.
195. Bacharach E, Itin A, Keshet E. Apposition-dependent
induction of plasminogen activator inhibitor type 1 expression:
a mechanism for balancing pericellular proteolysis during
angiogenesis. Blood 1998;92:939945.
196. Dublin E, Hanby A, Patel NK, et al. Immunohistochemical
expression of uPA, uPAR, and PAI-1 in breast carcinoma.
Fibroblastic expression has strong associations with tumor
pathology. Am J Pathol 2000;157: 12191227.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

57 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

197. Pedersen AN, Christensen IJ, Stephens RW, et al. The


complex between urokinase and its type-1 inhibitor in primary
breast cancer: relation to survival. Cancer Res
2000;60:69276934.
198. Soff GA, Sanderowitz J, Gately S, et al. Expression of
plasminogen activator inhibitor type I by human prostate
carcinoma cells inhibits primary tumor growth,
tumor-associated angiogenesis, and metastasis to lung and liver
in any athymic mouse model. J Clin Invest 1995;96:25932600.
199. Eitzman DT, Krauss JC, Shen T, et al. Lack of plasminogen
activator inhibitor-1 effect in a transgenic mouse model of
metastatic melanoma. Blood 1996;87:47184722.
200. Jankun J, Keck RW, Skrzypczak-Jankun E, et al. Inhibitors
of urokinase reduce size of prostate cancer xenografts in severe
combined immunodeficient mice. Cancer Res 1997;57:559563.
201. Bajou K, Noel A, Gerard RD, et al. Absence of host
plasminogen activator inhibitor 1 prevents cancer invasion and
vascularization. Nat Med 1998; 4:923928.
202. Gutierrez LS, Schulman A, Brito-Robinson T, et al. Tumor
development is retarded in mice lacking the gene for urokinasetype plasminogen activator or its inhibitor, plasminogen
activator inhibitor-1. Cancer Res 2000; 60:58395847.
203. Bajou K, Masson V, Gerard RD, et al. The plasminogen
activator inhibitor PAI-1 controls in vivo tumor vascularization
by interaction with proteases, not vitronectin. Implications for
antiangiogenic strategies. J Cell Biol 2001;152:777784.
204. McMahon GA, Petitclerc E, Stefansson S, et al.
Plasminogen activator inhibitor-1 regulates tumor growth and
angiogenesis. J Biol Chem 2001; 276:3396433968.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

58 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

205. Devy L, Blacher S, Grignet-Debrus C, et al. The pro- or


antiangiogenic effect of plasminogen activator inhibitor 1 is
dose dependent. FASEB J 2002;16:147154.
206. Curino A, Mitola DJ, Aaronson H, et al. Plasminogen
promotes sarcoma growth and suppresses the accumulation of
tumor-infiltrating macrophages. Oncogene 2002;21:88308842.
207. Vu TH, Hung DT, Wheaton VI, et al. Molecular cloning of a
functional thrombin receptor reveals a novel proteolytic
mechanism of receptor activation. Cell 1991;64:10571068.
208. Erickson LA, Hekman CM, Loskutoff DJ. The primary
plasminogen-activator inhibitors in endothelial cells, platelets,
serum, and plasma are immunologically related. Proc Natl Acad
Sci U S A 1985;82:87108714.
209. Booth NA, Simpson AJ, Croll A, et al. Plasminogen
activator inhibitor (PAI-1) in plasma and platelets. Br J
Haematol 1988;70:327333.
210. Declerck PJ, Verstreken M, Kruithof EKO, et al.
Measurement of plasminogen activator inhibitor 1 in biologic
fluids with a murine monoclonal antibody-based enzyme-linked
immunoabsorbent assay. Blood 1988;71: 220225.
211. Fay WP, Eitzman DT, Shapiro AD, et al. Platelets inhibit
fibrinolysis in vitro by both plasminogen activator inhibitor-1
dependent and independent mechanisms. Blood
1994;83:351356.
212. Fay WP, Parker AC, Condrey LR, et al. Human plasminogen
activator inhibitor-1 (PAI-1) deficiency: characterization of a
large kindred with a null mutation in the PAI-1 gene. Blood
1997;90:204208.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

59 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

213. Fay WP, Shapiro AD, Shih JL, et al. Complete deficiency of
plasminogen-activator inhibitor type 1 due to a frame-shift
mutation. N Engl J Med 1992;327:17291733.
214. Minowa H, Takahashi Y, Tanaka T, et al. Four cases of
bleeding diathesis in children due to congenital plasminogen
activator inhibitor-1 deficiency. Haemostasis 1999;29:286291.
215. Carmeliet P, Moons L, Lijnen R, et al. Inhibitory role of
plasminogen activator inhibitor-1 in arterial wound healing and
neointima formation: a gene targeting and gene transfer study
in mice. Circulation 1997;96: 31803191.
216. Carmeliet P, Moons L, Herbert JM, et al. Urokinase but not
tissue plasminogen activator mediates arterial neointima
formation in mice. Circ Res 1997;81:829839.
217. Redmond EM, Cullen JP, Cahill PA, et al. Endothelial cells
inhibit flow-induced smooth muscle cell migration: role of
plasminogen activator inhibitor-1. Circulation
2001;103:597603.
218. Ploplis VA, Cornelissen I, Sandoval-Cooper MJ, et al.
Remodeling of the vessel wall after copper-induced injury is
highly attenuated in mice with a total deficiency of plasminogen
activator inhibitor-1. Am J Pathol 2001; 158:107117.
219. Ploplis VA, Castellino FJ. Attenuation of neointima
formation following arterial injury in PAI-1 deficient mice. Ann
NY Acad Sci 2001;936:466468.
220. Zhu Y, Farrehi PM, Fay WP. Plasminogen activator inhibitor
type 1 enhances neointima formation after oxidative vascular
injury in atherosclerosis-prone mice. Circulation
2001;103:31053110.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

60 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

221. Peng L, Bhatia N, Parker AC, et al. Endogenous vitronectin


and plasminogen activator inhibitor-1 promote neointima
formation in murine carotid arteries. Arterioscler Thromb Vasc
Biol 2002;22:934939.
222. DeYoung MB, Tom C, Dichek DA. Plasminogen activator
inhibitor type 1 increases neointima formation in
balloon-injured rat carotid arteries. Circulation
2001;104:19721971.
223. Juhan-Vague I, Collen D. On the role of coagulation and
fibrinolysis in atherosclerosis. Ann Epidemiol 1992;2:427438.
224. Schneiderman J, Sawdey MS, Keeton MR, et al. Increased
type 1 plasminogen activator inhibitor gene expression in
atherosclerotic human arteries. Proc Natl Acad Sci U S A
1992;89:69987002.
225. Lupu F, Bergonzelli GE, Heim DA, et al. Localization and
production of plasminogen activator inhibitor-1 in human
healthy and atherosclerotic arteries. Arterioscler Thromb
1993;13:10901100.
226. Robbie LA, Booth NA, Brown AJ, et al. Inhibitors of
fibrinolysis are elevated in atherosclerotic plaque. Arterioscler
Thromb Vasc Biol 1996;16: 539545.
227. Sjoland H, Eitzman DT, Gordon D, et al. Atherosclerosis
progression in LDL receptor-deficient and apolipoprotein
E-deficient mice is independent of genetic alterations in
plasminogen activator inhibitor-1. Arterioscler Thromb Vasc Biol
2000;20:846852.
228. Eitzman DT, Westrick RJ, Xu Z, et al. Plasminogen
activator inhibitor-1 deficiency protects against atherosclerosis
progression in the mouse carotid artery. Blood

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

61 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

2000;96:42124215.
229. Eren M, Painter CA, Atkinson JB, et al. Age-dependent
spontaneous coronary arterial thrombosis in transgenic mice
that express a stable form of human plasminogen activator
inhibitor-1. Circulation 2002;106:491496.
230. Carmeliet P, Kieckens L, Schoonjans L, et al. Plasminogen
activator inhibitor-1 gene-deficient mice. I. Generation by
homologous recombination and characterization. J Clin Invest
1993;92:27462755.
231. Yamamoto K, Takeshita K, Shimokawa T, et al.
Plasminogen activator inhibitor-1 is a major stress-regulated
gene: implications for stress-induced thrombosis in aged
individuals. Proc Natl Acad Sci U S A 2002;99:890895.
232. Juhan-Vague I, Moerman B, De Cock F, et al. Plasma levels
of a specific inhibitor of tissue-type plasminogen activator (and
urokinase) in normal and pathological conditions. Thromb Res
1984;33:523530.
233. Meade TW, Ruddock V, Stirling Y, et al. Fibrinolytic
activity, clotting factors, and long-term incidence of ischaemic
heart disease in the Northwick Park Heart Study. Lancet
1993;342:10761079.
234. Thompson SG, Kienast J, Pyke SD, et al. Hemostatic
factors and the risk of myocardial infarction or sudden death in
patients with angina pectoris. European concerted action on
thrombosis and disabilities angina pectoris study group. N Engl
J Med 1995;332:635641.
235. Folsom AR, Aleksic N, Park E, et al. Prospective study of
fibrinolytic factors and incident coronary heart disease: the
Atherosclerosis Risk in Communities (ARIC) study. Arterioscler

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

62 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

Thromb Vasc Biol 2001;21:611617.


236. Alessi MC, Peiretti F, Morange P, et al. Production of
plasminogen activator inhibitor 1 by human adipose tissue:
possible link between visceral fat accumulation and vascular
disease. Diabetes 1997;46:860867.
237. Schafer K, Fujisawa K, Konstantinides S, et al. Disruption
of the plasminogen activator inhibitor 1 gene reduces the
adiposity and improves the metabolic profile of genetically
obese and diabetic ob/ob mice. FASEB J 2001;15:18401842.
238. Eriksson P, Kallin B, 't Hooft FM, et al. Allele-specific
increase in basal transcription of the plasminogen-activator
inhibitor 1 gene is associated with myocardial infarction. Proc
Natl Acad Sci U S A 1995;92:18511855.
239. Ridker PM, Hennekens CH, Lindpaintner K, et al. Arterial
and venous thrombosis is not associated with the 4G/5G
polymorphism in the promoter of the plasminogen activator
inhibitor gene in a large cohort of US men. Circulation
1997;95:5962.
240. Zoller B, Garcia dF, Dahlback B. A common 4G allele in the
promoter of the plasminogen activator inhibitor-1 (PAI-1) gene
as a risk factor for pulmonary embolism and arterial thrombosis
in hereditary protein S deficiency. Thromb Haemost
1998;79:802807.
P.380
241. Zhang ZG, Chopp M, Goussev A, et al. Cerebral
microvascular obstruction by fibrin is associated with
upregulation of PAI-1 acutely after onset of focal embolic
ischemia in rats. J Neurosci 1999;19:1089810907.
242. Hosomi N, Lucero J, Heo JH, et al. Rapid differential

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

63 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

endogenous plasminogen activator expression after acute


middle cerebral artery occlusion. Stroke 2001;32:13411348.
243. Nagai N, De Mol M, Lijnen HR, et al. Role of plasminogen
system components in focal cerebral ischemic infarction: a gene
targeting and gene transfer study in mice. Circulation
1999;99:24402444.
244. Chen CH, Eng HL, Chang CJ, et al. 4G/5G promoter
polymorphism of plasminogen activator inhibitor-1, lipid
profiles, and ischemic stroke. J Lab Clin Med
2003;142:100105.
245. Morange PE, Aubert J, Peiretti F, et al. Glucocorticoids and
insulin promote plasminogen activator inhibitor 1 production by
human adipose tissue. Diabetes 1999;48:890895.
246. Sakamoto T, Woodcock-Mitchell J, Marutsuka K, et al.
TNF-alpha and insulin, alone and synergistically, induce
plasminogen activator inhibitor-1 expression in adipocytes. Am
J Physiol 1999;276:C1391C1397.
247. Samad F, Schneiderman J, Loskutoff D. Expression of
fibrinolytic genes in tissues from human atherosclerotic
aneurysms and from obese mice. Ann NY Acad Sci
1997;811:350358.
248. Lundgren CH, Brown SL, Nordt TK, et al. Elaboration of
type-1 plasminogen activator inhibitor from adipocytes. A
potential pathogenetic link between obesity and cardiovascular
disease. Circulation 1996;93:106110.
249. Cigolini M, Tonoli M, Borgato L, et al. Expression of
plasminogen activator inhibitor-1 in human adipose tissue: a
role for TNF-alpha? Atherosclerosis 1999;143:8190.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

64 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

250. Crandall DL, Groeling TM, Busler DE, et al. Release of


PAI-1 by human preadipocytes and adipocytes independent of
insulin and IGF-1. Biochem Biophys Res Commun
2000;279:984988.
251. Gottschling-Zeller H, Rohrig K, Hauner H. Troglitazone
reduces plasminogen activator inhibitor-1 expression and
secretion in cultured human adipocytes. Diabetologia
2000;43:377383.
252. Vague P, Juhan-Vague I, Aillaud MF, et al. Correlation
between blood fibrinolytic activity, plasminogen activator
inhibitor level, plasma insulin level, and relative body weight in
normal and obese subjects. Metabolism 1986;35:250253.
253. Mavri A, Stegnar M, Krebs M, et al. Impact of adipose
tissue on plasma plasminogen activator inhibitor-1 in dieting
obese women. Arterioscler Thromb Vasc Biol
1999;19:15821587.
254. Giltay EJ, Elbers JM, Gooren LJ, et al. Visceral fat
accumulation is an important determinant of PAI-1 levels in
young, nonobese men and women: modulation by cross-sex
hormone administration. Arterioscler Thromb Vasc Biol
1998;18:17161722.
255. Samad F, Loskutoff DJ. Tissue distribution and regulation
of plasminogen activator inhibitor-1 in obese mice. Mol Med
1996;2:568582.
256. Shimomura I, Funahashi T, Takahashi M, et al. Enhanced
expression of PAI-1 in visceral fat: possible contributor to
vascular disease in obesity. Nat Med 1996;2:800803.
257. Ma LJ, Mao SL, Taylor KL, et al. Prevention of obesity and
insulin resistance in mice lacking plasminogen activator

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

65 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

inhibitor 1. Diabetes 2004; 53:336346.


258. Calles-Escandon J, Ballor D, Harvey-Berino J, et al.
Amelioration of the inhibition of fibrinolysis in elderly, obese
subjects by moderate energy intake restriction. Am J Clin Nutr
1996;64:711.
259. Samad F, Uysal KT, Wiesbrock SM, et al. Tumor necrosis
factor alpha is a key component in the obesity-linked elevation
of plasminogen activator inhibitor 1. Proc Natl Acad Sci U S A
1999;96:69026907.
260. Duymelinck C, Dauwe SE, De Greef KE, et al. TIMP-1 gene
expression and PAI-1 antigen after unilateral ureteral
obstruction in the adult male rat. Kidney Int
2000;58:11861201.
261. Eddy AA, Giachelli CM. Renal expression of genes that
promote interstitial inflammation and fibrosis in rats with
protein-overload proteinuria. Kidney Int 1995;47:15461557.
262. Brown NJ, Nakamura S, Ma L, et al. Aldosterone modulates
plasminogen activator inhibitor-1 and glomerulosclerosis in
vivo. Kidney Int 2000; 58:12191227.
263. Keeton M, Ahn C, Eguchi Y, et al. Expression of type 1
plasminogen activator inhibitor in renal tissue in murine lupus
nephritis. Kidney Int 1995; 47:148157.
264. Rerolle JP, Hertig A, Nguyen G, et al. Plasminogen
activator inhibitor type 1 is a potential target in renal
fibrogenesis. Kidney Int 2000;58:18411850.
265. Hamano K, Iwano M, Akai Y, et al. Expression of
glomerular plasminogen activator inhibitor type 1 in
glomerulonephritis. Am J Kidney Dis 2002;39:695705.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

66 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

266. Xu Y, Hagege J, Mougenot B, et al. Different expression of


the plasminogen activation system in renal thrombotic
microangiopathy and the normal human kidney. Kidney Int
1996;50:20112019.
267. Bergstein JM, Riley M, Bang NU. Role of
plasminogen-activator inhibitor type 1 in the pathogenesis and
outcome of the hemolytic uremic syndrome. N Engl J Med
1992;327:755759.
268. Nakamura T, Tanaka N, Higuma N, et al. The localization
of plasminogen activator inhibitor-1 in glomerular subepithelial
deposits in membranous nephropathy. J Am Soc Nephrol
1996;7:24342444.
269. Border WA, Noble NA. Transforming growth factor beta in
tissue fibrosis. N Engl J Med 1994;331:12861292.
270. Border WA, Okuda S, Nakamura T. Extracellular matrix and
glomerular disease. Semin Nephrol 1989;9:307317.
271. Lund LR, Riccio A, Andreasen PA, et al. Transforming
growth factor-beta is a strong and fast acting positive regulator
of the level of type-1 plasminogen activator inhibitor mRNA in
WI-38 human lung fibroblasts. EMBO J 1987;6:12811286.
272. Kerins DM, Hao Q, Vaughan DE. Angiotensin induction of
PAI-1 expression in endothelial cells is mediated by the
hexapeptide angiotensin IV. J Clin Invest 1995;96:25152520.
273. Oikawa T, Freeman M, Lo W, et al. Modulation of
plasminogen activator inhibitor-1 in vivo: a new mechanism for
the anti-fibrotic effect of renin-angiotensin inhibition. Kidney
Int 1997;51:164172.

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

67 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

274. Huang Y, Haraguchi M, Lawrence DA, et al. A mutant,


noninhibitory plasminogen activator inhibitor type 1 decreases
matrix accumulation in experimental glomerulonephritis. J Clin
Invest 2003;112:379388.
275. Fogo AB. Renal fibrosis: not just PAI-1 in the sky. J Clin
Invest 2003; 112:326328.
276. Ingelfinger JR. Forestalling fibrosis. N Engl J Med
2003;349:22652266.
277. Bertozzi P, Astedt B, Zenzius L, et al. Depressed
bronchoalveolar urokinase activity in patients with adult
respiratory distress syndrome. N Engl J Med
1990;322:890897.
278. Idell S, Peters J, James KK, et al. Local abnormalities of
coagulation and fibrinolytic pathways that promote alveolar
fibrin deposition in the lungs of baboons with diffuse alveolar
damage. J Clin Invest 1989;84:181193.
279. Chapman HA, Allen CL, Stone OL. Abnormalities in
pathways of alveolar fibrin turnover among patients with
interstitial lung disease. Am Rev Respir Dis 1986;133:437443.
280. Kotani I, Sato A, Hayakawa H, et al. Increased
procoagulant and antifibrinolytic activities in the lungs with
idiopathic pulmonary fibrosis. Thromb Res 1995;77:493504.
281. Hasday JD, Bachwich PR, Lynch JP, et al. III. Procoagulant
and plasminogen activator activities of bronchoalveolar fluid in
patients with pulmonary sarcoidosis. Exp Lung Res
1988;14:261278.
282. Barazzone C, Belin D, Piguet PF, et al. Plasminogen

24/06/2006 01:13 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

68 de 68

http://gateway.ut.ovid.com/gw1/ovidweb.cgi

activator inhibitor-1 in acute hyperoxic mouse lung injury. J


Clin Invest 1996;98:26662673.
283. Viscardi RM, Broderick K, Sun CC, et al. Disordered
pathways of fibrin turnover in lung lavage of premature infants
with respiratory distress syndrome. Am Rev Respir Dis
1992;146:492499.
284. Idell S, Gonzalez K, Bradford H, et al. Procoagulant
activity in bronchoalveolar lavage in the adult respiratory
distress syndrome. Contribution of tissue factor associated with
factor VII. Am Rev Respir Dis 1987; 136:14661474.
285. Olman MA, Mackman N, Gladson CL, et al. Changes in
procoagulant and fibrinolytic gene expression during
bleomycin-induced lung injury in the mouse. J Clin Invest
1995;96:16211630.
286. Eitzman DT, McCoy RD, Zheng X, et al. Bleomycin-induced
pulmonary fibrosis in transgenic mice that either lack or
overexpress the murine plasminogen activator inhibitor-1 gene.
J Clin Invest 1996;97:232237.
287. Aertgeerts K, De Bondt HL, De Ranter CJ, et al.
Mechanisms contributing to the conformational and functional
flexibility of plasminogen activator inhibitor-1. Nat Struct Biol
1995;2:891897.

24/06/2006 01:13 p.m.

Vous aimerez peut-être aussi