You are on page 1of 138

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

1 de 138

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 18 Pl as m i nogenP l asm i n System

Chapter 18
PlasminogenPlasmin System
Nuala A. Booth
Fedor Bachmann
Hemostasis requires mechanisms both to stop bleeding by
formation of a hemostatic plug and to limit the plug's development,
allowing reestablishment of normal blood flow. The latter function
is largely accomplished by localized activation of the
plasminogenplasmin enzyme system, also called the fibrinolytic
system. The system is finely tuned, as it has to be to accomplish
healing of a vascular lesion without compromising the early
stability of the hemostatic plug and to limit activity to the injured
area. There is a dynamic balance between coagulant and
fibrinolytic activities, each of which represents a further balance
between proteolytic and inhibitory proteins. Excessive local or
systemic fibrinolytic activity can result in bleeding, often occurring
after a delay, as the weakened plug is dissolved. Conversely, an
inadequate fibrinolytic response may retard lysis of a thrombus
and contribute to its extension.
Plasmin is the main fibrinolytic enzyme. It is a trypsinlike serine
protease that degrades fibrin, and it is generated by activation of
the zymogen, plasminogen. The central players dictating the
activity of the system are the plasminogen activators (PAs) and the
inhibitors of both the activation stage and of plasmin activity. The

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

2 de 138

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

two major PAs that occur in the circulating blood, tissue-type


plasminogen activator (tPA) and urinary-type plasminogen
activator (uPA), also called urokinase or UK, are also serine
proteases. Inhibitors, most of which are of the serine protease
inhibitor (serpin) family, regulate the proteases. These are PA
inhibitors, plasminogen activator inhibitor 1 (PAI-1; see Chapter
19) and PAI-2, and 2 -antiplasmin the principal inhibitor of
plasmin. Receptors and binding proteins for plasminogen and the
PAs on endothelial cells (see Chapter 22), platelets and leucocytes
regulate local generation of plasmin. The best characterized of
these receptors is the urinary-type plasminogen activator receptor
(uPAR). Tables 18-1 and 18-2 list several of the properties of the
genes, messenger ribonucleic acids (mRNAs), and proteins of the
components of the fibrinolytic system, together with modulating
factors. These include factor XII and the contact phase of
coagulation, in which interactions with prekallikrein and kininogen
also promote plasminogen activation. They also include
apolipoprotein (a), cytokines, growth factors and hormones (see
Chapter 6), the PAI-1binding protein vitronectin, the
thrombin-activatable fibrinolytic inhibitor (TAFI) (see Chapter 20),
the low density lipoprotein receptorrelated protein (LRP, identical
to the 2 -macroglobulin receptor), the mannose receptor, annexin
II, and several other inhibitors that affect the physiology of
fibrinolysis (see Chapter 23).
Like the serine proteases of blood coagulation, fibrinolytic
proteases generally exist in a zymogen form, a single-chain form
of the molecule that is activated by cleavage of a single peptide
bond. All the active serine proteases cleave arginine and/or lysine
bonds and originated from a simple trypsinlike ancestor protease
(1,2). In the case of the fibrinolytic proteases, the products are all
in a two-chain form held together by one or more disulfide bonds.
An exception to this rule is tPA, which is an active protease in its
single-chain form; tPA can be considered to be at one end of the
scale of zymogenicity, with plasminogen as a totally inactive
example at the other (3).
The typical structure of these proteases (Fig. 18-1) is that they
have an N-terminal A chain containing several independently folded
modules or domains, which have particular binding properties, and

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

3 de 138

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

a C-terminal B chain, which is the protease domain. The kringle


domainnamed after a Danish pastryis approximately 80
residues long and is held together by three internal disulfide
bonds. The finger domains are homologous to type I and type II
fibronectin regions. Type I finger domains are approximately 50
residues long, contain two internal disulfide bonds, and may play a
role in fibrin binding. The epidermal growth factor (EGF) domains,
which are homologous with a module in the EGF precursor, contain
53 residues and three internal disulfide bonds. In several proteins,
these domains bind to specific cell-surface receptors or binding
proteins. The apple domain has been found only in prekallikrein
and factor XI, in which it occurs as a quadruple repeat of 90
residues in the N-terminal portion of these molecules. Like the
kringle domains, it contains three internal disulfide bonds. The
protease domain contains the catalytic triad, Ser195, His57, and
Asp102 (chymotrypsin numbering), and the domain comprises
approximately 250 amino acid residues and four disulfide bonds.
The protease domain is well conserved, with particularly high
sequence identity in the vicinity of the three active site residues
and at the N-terminus.
The components of the fibrinolytic system are discussed next,
beginning with the central enzyme, plasmin, and its precursor,
plasminogen, then considering in turn plasminogen activators, the
different inhibitors of plasmin, and plasmin generation, with a final
summary on the balance of the system.

PLASMINOGEN
Human plasminogen (PLG) is the zymogen form of the main
fibrinolytic protease, plasmin (EC 3.4.21.7). PLG is a single-chain
glycoprotein of 92 kDa, consisting of 791 amino acid residues and
approximately 2% carbohydrate (Table 18-1). In the human adult,
the plasma concentration of PLG is approximately 200 mg per L,
which corresponds to M (Table 18-2). The half-life (t 1/ 2 ) of native
PLG, which has a glutamic acid residue at its N-terminus
(GluPLG), is approximately 2.2 days; the slightly degraded
LysPLG (see subsequent section, Activation of Plasminogen to
Plasmin) has a much shorter t 1/ 2 of 0.8 days (4). The principal
production site for PLG is the liver, but PLG is present in the

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

4 de 138

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

extravascular space of most tissues, and some, such as


eosinophils, the kidney, and the cornea, may be capable of
synthesizing it (5,6,7,8).

FIGURE 18-1. Structural elements of some serine proteases.


A, apple domain (approximately 80 residues); AP, activation
peptide; E, epidermal growth factor domain (approximately 32
residues); F, fibronectin finger (approximately 40 residues); K,
kringle (approximately 90 residues); P, protease domain (B
chain) (approximately 250 residues); scuPA, single-chain
urokinase; tPA, tissue-type plasminogen activator.

TABLE 18-1 CHROMOSOME LOCATION AND GENE ORGANIZATIO


CONSTITUENTS OF THE PLASMINOGEN-PLASMIN SYSTEM

Factor

Symbol

Chromosome

Gene

mRNA

(kb)

(kb)

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

5 de 138

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

Plasminogen

PLG

6q26-q27

52.5

2.9

tPA

PLAT

8p12-p11

32.7

2.7

uPA

PLAU

10q24

6.4

2.4

PAI-1

PLANH1

7q21.3-q22

12.2

2.4/3.2

PAI-2

PLANH2

18q22.1

16.5

1.9

Factor XII

F12

5q33-qter

12

2.6

Prekallikrein

KLK3

4q34-q35

22

2.4

HMW kininogen

KNG

3q27

27

3.2

Vitronectin

VTN

17q11

4.5

1.6

C1-inhibitor

C1NH

11q12-q13.1

17

1.8

2 -Antiplasmin

PLI

17p13

16

2.2

2 -Macroglobulin

A2M

12p13.3-p12.3

48

4.6

Histidine-rich

HRG

3q27

15.5

2.1

TNA

3p22-p21.3

12

0.9

APOAL2.1

6q26-27

glycoprotein
Tetranectin,
monomer
Apolipoprotein

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

6 de 138

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

(a)
TAFI

CPB

13q14.11

48

1.8

uPAR

PLAUR

19q13.1-q13.2

23

1.4

Annexin II a

ANX2

15q21-q22

1.3

LRP ( 2 -MR)

LRP1

12q13-q13.3

92

15

-Enolase

ENO1

1pter-1p36.13

>18

1.7

Mannose-receptor

MRC1

10p13

5.1

mRNA, messenger ribonucleic acid; tPA, tissue-type plasminogen activ


uPA, urinary-type plasminogen activator; PAI-1, plasminogen activato
inhibitor 1; PAI-2, plasminogen activator inhibitor 2; HMW,
high-molecular-weight; C1-inhibitor, complement component 1 esteras
inhibitor; TAFI, thrombin-activatable fibrinolysis inhibitor; uPAR,
urinary-type plasminogen activator receptor; LRP, low density lipoprot
receptorrelated protein; MR, macroglobulin receptor.
a

Heavy chain of tetramer.

Variable, depending on the number of kringle 4 inserts.

TABLE 18-2 PROPERTIES OF PROTEIN CONSTITUENTS OF TH

Amino
Factor
Plasminogen

M r (kDa)

acid
residues

Concentration
(mg/L)

92

791

200

Mo
concent
2

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

7 de 138

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

tPA

68

530

0.005

70

uPA

54

411

0.002

40

PAI-1

52

379

0.01

200

PAI-2

46/70

393

<0.005

<70

Factor XII

80

596

30

0.375

Prekallikrein

88

619

40

0.45

110

626

70

0.60

Vitronectin

78

459

350

4.5

C1-inhibitor

105

478

180

1.7

70

452

70

725

1,451

2,500

75

507

100

HMW kininogen

2 -Antiplasmin
2 -Macroglobulin
Histidine-rich

1.5

glycoprotein

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

8 de 138

Tetranectin,

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

22.5

181

10

0.5

300800

<7

TAFI (pro-CpU)

60

401

uPAR

55

313

Annexin II b , f

38

338

LRP ( 2 -MR)

600

4,525

54

433

175

1,437

monomer b , c
Apolipoprotein(a) d

-Enolase
Mannose receptor

75

z, protease zymogen; tPA, tissue-type plasminogen activator; p, prote


activator; PAI-1, plasminogen activator inhibitor 1; i, inhibitor; PAI-2,
high-molecular-weight; c, cofactor; C1-inhibitor, complement compone
TAFI, thrombin-activatable fibrinolysis inhibitor; uPAR, urinary-type p
density lipoprotein receptorrelated protein; MR, macroglobulin recept
M r of mature proteins is listed.
a

Estimated value, calculated from difference of molecular weight deter

amino acids derived from complementary DNA.


b
c

M r of nonglycosylated form as listed in Swiss protein base ExPASy.

Heavy chain of homotrimer.

Many isoforms due to 13 up to 37 type 4 kringle motifs.

Activated form.

Heavy chain of heterotetramer (formed by two heavy and two 10-kDa

P.336

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

9 de 138

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

P.337

Structural Organization
The PLG gene is 52.5 kb, the largest among the fibrinolytic
proteins. It is located on chromosome 6q26-q27 (9), close to two
genes for apolipoprotein (a) and for the PLG-related genes A and B
(10,11). Amino acid sequence analysis showed the structural
features of N-terminal activation peptide, five kringles, and the
protease domain (12). Subsequent isolation of mRNA, and later of
the gene, confirmed the original sequence, except for an additional
isoleucine residue at position 67 (13).
The 19 exons of the PLG gene are separated by 18 introns that
follow the general GT-AG rule found in other eukaryotic genes. The
first exon encodes the signal sequence, with two exons encoding
the activation peptide and each kringle; the remaining six exons
code for the intervening sequence and the protease domain (see
Fig. 18-2). Each of the five kringles consists of 78 to 80 residues,
and they are well-conserved between species (14). Functionally,
the kringles give PLG the ability to bind to exposed lysyl residues
in fibrin (15,16), -antiplasmin (17), tetranectin (18),
histidine-rich glycoprotein (19), collagen (20),
high-molecular-weight (HMW) and low-molecular-weight (LMW)
kininogen (21), thrombospondin (22), and cell surface receptors.
The affinity of the kringles for lysyl residues is also exploited to
isolate it from human plasma by affinity chromatography (23).
The kringle structure is shaped by a characteristic 16, 24, 35
disulfide bond pattern involving six conserved Cys residues in
positions 1, 22, 51, 63, 75, and 80 (kringle 5 numbering). A single
nonconserved Cys in position 4 of kringle 2 forms an interkringle
disulfide bond with Cys43 in kringle 3 (not shown on Fig. 18-2)
and thereby restricts the mobility of kringles 2 and 3. Solution
structures have been determined for kringle 1 (24), kringle 2 (25),
kringle 2 + 3 (26), and kringle 4 (27). Crystallographic structures
have been solved for kringle 1 (28,29), kringle 4 (30), and kringle
5 (31). The human PLG kringles have different affinities and
specificities for -amino acid ligands. The tightest binding site for
-aminocaproic acid (ACA) is provided by kringle 1 [dissociation
constant (K d ) of 9M] (32,33) followed by kringle 4 (32,34,35) and

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

10 de 138

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

kringle 5 (34,35). Kringle 2 displays only a weak interaction with


ACA probably not of functional significance, whereas kringle 3
shows no interaction (36).
Experimentally, kringle 4 can be isolated from PLG by digestion
with elastase, which cleaves after small amino acids and at
relatively frequent ValVal sequences. Therefore, digestion with
elastase typically yields three major components: a fragment
containing kringles 1 through 3, isolated kringle 4, and kringle 5
attached to the protease portion of PLG (12). The latter fragment,
also called mini-PLG, can be activated to plasmin by PAs
(37,38,39). Although it has some affinity for fibrin (16,40), it does
not bind to lysineSepharose and therefore is separated from
kringles 13 and kringle 4, which do bind lysine (12,32). Kringle 1
can be obtained from kringles 13 by further digestion with other
enzymes, such as Staphylococcus aureus V8 protease (41).

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

11 de 138

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

FIGURE 18-2. Sequence of the human plasminogen molecule.


Positions of the 18 introns are indicated by solid arrows at (for
type 1 and 2 introns) or in between (for type 0 introns)
residues. The 19-residue signal peptide (shown in a box with
negative numbers) is cleaved by signal peptidase at the
GlyGlu peptide bond. Conversion of plasminogen to plasmin
occurs after the peptide bond Arg561Val562 (shown by an
open, curved arrow) is cleaved by plasminogen activators. PAP
refers to the preactivation peptide generated by the action of
plasmin on Gluplasminogen; primary (but not sole) cleavage
site is Lys77Lys78 bond (shown by an open, straight arrow).
K1K5, kringles, located in the A chain; the B chain remains
attached to the A chain after cleavage of the 561562 bond via
the disulfide bonds Cys548Cys666 and Cys558Cys566.
Carbohydrate attachment sites (Asn289 and Thr346) are shown
by diamonds. (From Petersen TE, Martzen MR, Ichinose A, et
al. Characterization of the gene for human plasminogen, a key
proenzyme in the fibrinolytic system. J Biol Chem
1990;265:6104, with permission.)

P.338
Crystallography has revealed that the kringle 4 structure is highly
stabilized by an internal hydrophobic core and an extensive
hydrogen bonding network. The lysine binding site (LBS) on the
surface of kringle 4 is an open, elongated, shallow trough, formed
by the hydrophobic residues Trp62, Phe64, and Trp72, surrounded
by the positively charged Lys35 and Arg71 on one side and, at a
distance of approximately 7 , the negatively charged side chains
of Asp55 and Asp57 (see Fig. 18-3) (30). This structure provides
for ideal docking of zwitterions such as lysine or ACA, whose
opposite charges also are approximately 7 apart. The interaction
between binding site and ligand is enhanced further by the close
van der Waals contacts between the methylene carbons of ligand
and the aromatic residues of the amino acids in the center of the
binding site (30). The binding sites in kringles 1 and 5 are
similarly constructed. In kringle 1, Arg34 takes the place of Lys35
in kringle 4 (30,42); in kringle 5, Leu71 is substituted for Arg71 in

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

12 de 138

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

kringle 4 (31).
The catalytic domain of PLG (B chain) is made up of 230 residues
and is homologous with the trypsin family of serine proteases. In
the active site, it contains three amino acids: His603, Asp646, and
Ser741. Two groups of investigators have elucidated the crystal
structure of micro-PLG mutants. The main observations, a
deformed catalytic triad and a blocked S1 specificity pocket, are
similar in the two crystal structures (43,44). This unusual
structural feature has only been observed in the crystal structure
of factor D of the complement system, in contrast to the preformed
catalytic triad observed in other trypsinlike structures.

Activation of Plasminogen to Plasmin


All known PAs cleave the Arg561Val562 bond in PLG (see Fig.
18-4). The B chain of the resulting two-chain Gluplasmin
P.339
molecule remains attached to the A chain by two disulfide bonds.
When these PAs are added to GluPLG in vitro, the generation of
trace amounts of plasmin results in cleavage of the N-terminus by
hydrolysis of one or several of the following bonds: Arg68Met69,
Lys77Lys78, or Lys78 Val79 (45,46). The final product obtained
is Lysplasmin (Fig. 18-4). Under physiologic conditions, the
plasmin-mediated conversion to LysPLG probably does not occur
in circulating blood (47) because free plasmin is inactivated rapidly
by 2 -antiplasmin. Studies with monoclonal antibodies that
recognize LysPLG but not GluPLG revealed no free LysPLG nor
bound Lysplasmin in normal human plasma and only minimal
increases of these forms in patients who received thrombolytic
therapy (48).

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

13 de 138

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

FIGURE 18-3. Model of the interaction site of plasminogen


kringles 1 and 4 with -aminocaproic acid (depicted in heavy
lines), a lysine analog. A hydrophobic core on the surface of an
elongated shallow trough is surrounded by two clusters of
positively and negatively charged amino acids.

Several -aminocarboxylic acids, such as 6-aminohexanoic acid


(ACA, 6-AHA), p-aminomethyl benzoic acid (p-AMBA), and
trans-4-amino-methyl-cyclohexane-carboxylic acid (trans-AMCA,
also known as tranexamic acid), inhibit the functional activity of
plasmin in vitro and in vivo by binding to the LBS of
plasmin(ogen), thereby inhibiting the binding of plasmin(ogen) to
fibrin (49,50). These compounds are therefore termed
antifibrinolytic. It should be noted that their effects depend
strongly on the concentration used; at micromolar to low millimolar
concentrations, these compounds actually increase the activation
rate. They do this by inducing a conformational change, making
GluPLG adopt a structure similar to that of LysPLG, which is
more easily activated to plasmin. The change reflects the loss of
interaction of the N-terminal preactivation peptide with LBS on the
kringles (51,52). It causes a marked decrease in the Michaelis

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

14 de 138

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

constant K M for activation with tPA or uPA, by approximately one


order of magnitude, down to the range of the physiologic
concentration of PLG in human plasma (53,54). In GluPLG, Lys50
binds to LBSs on kringles (51,52), making these LBS less available
for binding to C-terminal lysine in fibrin, PLG receptors, and other
proteins. Therefore LysPLG binds with higher affinity to fibrin
than does GluPLG.
Conformational differences between Glu and LysPLG have been
revealed by small-angle neutron scattering. GluPLG goes from
having a radius of gyration of 39 to 56 in the presence of ACA
(55). The closed form, in which the five kringle domains and the
catalytic domain interact to produce an ellipsoid shape, is less
activatable than the open form (see Fig. 18-5),
P.340
whose increased flexibility facilitates the binding of PAs, leading to
approximately a 10-fold decrease of K M Furthermore, the open
form, taken up by PLG after binding to fibrin, enables the
proteolytic domain to sweep over a larger area and therefore be
more efficient in the digestion of fibrin than a closed form with
restricted mobility. Activatability of GluPLG by PAs is dependent
on other factors as well, such as the presence of anions, which in
the human plasma are mainly chloride ions, and of the divalent
cations Ca 2+ , Mg + and Mn + All these ions counteract the effect of
ACA and stabilize PLG in the Glu-form, which is poorly activatable
by PAs (54,56).

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

15 de 138

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

FIGURE 18-4. Activation of native Gluplasminogen. In the


absence of plasmin inhibitors, tissue plasminogen activator
(tPA) or urinary plasminogen activator (uPA) converts
Gluplasminogen to Glu plasmin; the latter then converts
Gluplasmin(ogen) to Lysplasmin(ogen).

FIGURE 18-5. A sketch of the conformational change in

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

16 de 138

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

plasminogen induced on occupation of a weak lysine binding


site by the ligand -aminocaproic acid. In the absence of
ligand, the domains in plasminogen interact to form a molecule
with the overall shape of a prolate ellipsoid. Binding of the
ligand converts this into an extended flexible structure in
which the interaction between the domains is abolished. The
closed and open forms are drawn to scale. (From Mangel WF,
Lin B, Ramakrishnan V. Characterization of an extremely large,
ligand-induced conformational change in plasminogen. Science
1990;248:69, with permission.)

Many modulators increase or decrease the rate of PLG activation;


the most important of these is fibrin. PLG binds to fibrin, directing
the process of clot lysis to its target, binding that is competed by
-amino carboxylic acids (57). Identification of the binding sites
was achieved by examining the products of elastase digestion of
PLG. Kringle 13 and miniplasminogen bound to fibrin, but kringle
4 did not (15,16,58,59). Notably, there is no close correlation
between the affinities of kringles for fibrin and lysine/ACA (see
Table 18-3).
Partial degradation of fibrin by plasmin creates a positive feedback
mechanism, whereby native GluPLG is bound to newly created
C-terminal lysines on the fibrin molecules, particularly on the
C-terminal portion of the (60,61,62,63,64). A few studies
support the concept that PLG might bridge adjoining fibrin
molecules and that probably both kringle 5 and kringle 1 bind to
fibrin, albeit not at the same binding site (59,65,66). There is little
consensus on the values of the dissociation constants and number
of binding sites for the binding of PLG, particularly of GluPLG to
fibrin. Taking all reported figures together, it appears that an
average K D for the binding of GluPLG to native fibrin is
approximately 5 M and that of LysPLG, roughly one order of
magnitude lower. In the presence of partially digested fibrin, the
K d for the binding of GluPLG approaches that of LysPLG for
native fibrin (0.5 M) (63,64). This probably means that GluPLG,
after binding to fibrin, takes on the open configuration (Fig. 18-5)

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

17 de 138

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

and functionally behaves like LysPLG.

TABLE 18-3 AFFINITIES OF PLASMINOGEN KRINGLES FOR


2 -AMINOCAPROIC ACID AND FOR NATIVE FIBRIN

Kringle

Affinity for

Affinity for native

number(s)

ACA

fibrin

13

High

Moderate

Moderate

Very low a

Low

High

A nuclear magnetic resonance imaging study has shown

that kringle 4 can bind to partially digested fibrin, which


has many C-terminal lysyl sites exposed (58). This binding
was abolished by -aminocaproic acid (ACA).

Recombinant microplasminogen, consisting of residues 542791,


can be activated to microplasmin by all four common PAs: tPA,
uPA, streptokinase (SK), and staphylokinase (SAK) (39).
Autocatalytically produced microplasmin consisting of the last 31
amino acid residues of the A chain and B chain exhibited a similar
affinity for fibrin as regular plasmin but a markedly lower catalytic
rate constant (k cat = 0.0076 per second vs. 0.064 per second for
plasmin) (38).

Cell Binding of Plasminogen


PLG binds to different cell types and a number of binding proteins
have been identified on human and other cells (67), -enolase (68)
and annexin II (69) being the most studied. All are abundant
proteins, and generally the binding occurs through Lys residues.
Interestingly, PLG binding to the prion protein PrPsc distinguishes
between the normal and misfolded conformers (70). The

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

18 de 138

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

association of PLG with the various binding proteins is not always


of high affinity, and the fact that so many proteins can serve this
function precludes their being viewed strictly as PLG receptors. The
abundance of such binding sites does make it likely that this
phenomenon indeed potentiates local plasmin generation (71).

Angiostatins
Many primary tumors have been found to suppress the growth of
their remote metastases or of a second and different tumor. At
least in part, this effect is due to angiostatin, originally described
as a 38-kDa protein and identified as an internal proteolytic
fragment of PLG comprising kringles 14 (72). It appears that
tumor-mediated proteolysis is responsible for the generation of
angiostatin, through uPA-induced PLG activation and the opening
up the inter-kringle and other disulfide bonds by plasmin
reductases (73,74), which include known PLG-binding proteins that
act as plasmin reductases (67,75,76). Similarly, proteolytic
cleavage of PLG by macrophage elastase (MMP-12), or
stromelysin-1 (MMP-3), or a 24-kDa endopeptidase from a
Chrysobacterium species generated angiostatinlike fragments
(77,78).
The antiproliferative effect of PLG fragments is not restricted to
the originally described kringle 14 or 15 fragments. Recombinant
fragments containing kringles 13 (79,80,81), or even isolated
kringles 1, 2, 3, or 5, but not kringle 4, had antiproliferative
activity similar to the originally described larger fragments
(79,82,83). On the other hand, isolated kringle 4 was potent in
inhibiting endothelial cell migration (84). Endothelial cell
proliferation assays revealed that angiostatin induces apoptosis
(85) and also appears to have a role in the prevention of
atherosclerosis, in that it inhibits smooth muscle cell proliferation
and migration in vitro (86). In all of these assays, intact PLG was
inactive. The sum of these data suggests that various internal
fragments of PLG exert powerful biological effects and that
fragments of HMW kininogen (87) and of apolipoprotein (a) (88)
have similar effects.
P.341

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

19 de 138

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

Variants of Plasminogen
The PLG molecule is subject to considerable variation in the normal
population. In addition to the limited proteolysis and activation
processes already discussed, the two important sources of
variation are glycosylation and genetic polymorphism.

Glycosylation
Human GluPLG occurs in two variants that differ in glycosylation.
PLG variant 1 is diglycosylated, with N-linked carbohydrate moiety
of the mannose type with 10 to 11 monosaccharide units on
Asn289, and an O-linked moiety of three to four residues on
Thr346 (89). PLG 2, which accounts for a little more than half of
the PLG molecules, contains only the carbohydrate group at Thr346
(90). Both of these forms exhibit, on isoelectric focusing,
additional heterogeneity with respect to their sialic acid content
(45,46,91,92). Therefore, each of the two variants produces six
additional isoforms with pIs between 6.1 and 6.6. All 12 molecular
forms are present in plasma from single donors (4).
In the absence of fibrin, PLG 1 appears to be more easily activated
to plasmin than is PLG 2, but this difference is not observed in the
presence of fibrin. Variant 1 changes its conformation more easily
in the presence of fibrin or tranexamic acid than does variant 2
(93). Variant 2 was cleared more rapidly than variant 1 and bound
five times faster to a deendothelialized vessel surface (94,95,96).
A novel glycosylation site containing a trisaccharide on Ser249 has
been described (97), and another may exist at Ser339 (98); a
further serine residue, Ser578, can be phosphorylated (99).

Polymorphism
Isoelectric focusing of neuraminidase-treated plasma (which
removes sialic acid) revealed several bands of activatable PLG,
representing the genetic polymorphism of PLG alleles. The
nomenclature, adopted in 1984 (100), is that the two most
commonly observed variants found in all races are called PLG A (A
for acidic pI) and PLG B (B for basic pI). Recently, the molecular
basis of these common polymorphisms was reported: PLG A have
Asp453, and PLG B have Asn453 (101). Variants with intermediate

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

20 de 138

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

pI are designated as PLG M. Variants with more acidic pI than PLG


A receive, in addition to the letter A, a numerical suffix that
increases with decreasing pI of the variant. The variants with more
basic pI than PLG B receive a numerical suffix that increases with
increasing pI of the variant. Alleles are designated with an asterisk
(e.g., PLG*A3), and silent (null) PLG alleles are designated
PLG*Q0, with Q0 standing for quantitatively zero or nonexpressed.
The frequency of the alleles in 1,330 unrelated West Germans was
70.1% for PLG*A, 27.7% for PLG*B, 1.5% for PLG*A3, 0.6% for
PLG*R (R: the sum of all other rare PLG alleles), and 0.35% for
PLG*Q0 (100). In the Japanese population, the frequency of the
allele PLG*A is considerably higher (95%) than in the Western
population (102).
Since the first report in 1978 of a hereditary molecular abnormality
in PLG in a patient with recurrent thrombosis (103), several
hundred cases have been described with a deficiency of PLG
antigen and activity (type I), or a normal level of antigen but
reduced activity (type II, dysplasminogenemia). In most instances,
a thromboembolic event or the presence of ligneous conjunctivitis
has been the starting point for family studies. Only those cases in
which the molecular defect has been identified are listed in Table
18-4
(101,102,103,104,105,106,107,108,109,110,111,112,113,114);
updates can be found
P.342
on the OMIM (Online Mendelian Inheritance in Man) Web site. The
sum of all data does not prove conclusively that the heterozygous
presence of one mutant PLG gene results in an increased incidence
of thromboembolism (115,116). In healthy subjects, the Ala601Thr
mutation is found in 2.2% and 2.9%, respectively, of the Japanese
and Chinese Han populations (107,112). Similarly, Lys19 to Glu
occurs relatively frequently in association with
hypoplasminogenemia (104,117), and there is controversy over
whether it should be regarded as a polymorphism (118).
Homozygous deficiencies, however, are clearly associated with
ligneous conjunctivitis, caused primarily by Arg216 to His or
Trp597 to stop mutations (106), and replacement therapy with PLG
is effective (108).

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

21 de 138

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

TABLE 18-4 CONGENITAL PLASMINOGEN MUTATIONS

Residue
9

Mutation
Thr to

References

Other
names

Type

Clinica
manifestat

106

106

101

104

101

106

105

Ligneous

Asn
19

Lys to
Glu

128

Lys to
Pro

133

Cys to
stop

134

Arg to
Lys

212

Lys
deletion

216

Arg
toHis

355

Val to

conjunctivi
107

II

Nagoya

Thromboph

107

Nagoya-I

Thromboph

Phe
374

Val to
Phe

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

22 de 138

453

Asp to

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

101

108

106

108

Thromboph

Asn
460

Glu to
stop

513

Arg to
His

572

Ser to
Pro

(secretion
defect)

597

Trp to

105,106

Cys
597

Trp to

conjunctivi
104

stop
601

Ala to

Ala to

Ligneous
conjunctivi

107,112

II

Thr
620

Ligneous

Tochigi,

Thromboph

Osaska II
107

Thr

Nagoya-II,

Thromboph

Tochigi,
Kagoshima

675

Ala to

109,111

112,113

II

Thromboph

Thr
676

Asp to
Asn

693

Gly to

Osaka-II,

None

III
101

II

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

23 de 138

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

Arg
732

Gly to

114

II

Kanagawa-I

Arg

It is not yet clear why some mutations cause ligneous


conjunctivitis and why very low functional PLG levels can be
tolerated in some individuals, without thrombosis occurring (106).
PLG-deficient mice, depending on genetic background, also develop
ligneous conjunctivitis (119). These mice are predisposed to
severe thrombosis and deposit intravascular and extravascular
fibrin in many tissues (120,121). Growth is retarded, rectal
prolapse is frequent, and Plg - /- mice die prematurely, but
ovulation, embryonic development, and reproduction are normal
(122). Wound healing and tissue remodeling are impaired
(123,124), and response to an inflammatory stimulus is diminished
(125). After electric injury to the femoral artery, wound healing,
removal of necrotic debris, leukocyte infiltration, smooth muscle
cell immigration, and arterial neointima formation were greatly
delayed in Plg - /- mice (126), suggesting that PLG deficiency might
reduce the development of atherosclerosis. However, in Plg - /- mice
crossed with apolipoprotein Edeficient mice prone to develop early
atherosclerotic lesions, the spontaneous development of intimal
lesions was greatly accelerated (127). Plg - /- mice crossed with
fibrinogan

- /-

mice were restored to normality in many of these

cases, demonstrating the causative role of intravascular and


extravascular fibrin deposition (128,129).

Regulation of Expression
Plasma levels of PLG are generally rather stable, as discussed
further in the section, Balance of the System, but there is an
increase in the acute-phase response (130). Two sequence
elements of CTGGGA common to acute-phase reactant genes are at
position 76 to 81 and -553 to -558 (13). Three interleukin-6
(IL-6)responsive elements are at positions -830, -518 and +117

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

24 de 138

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

in the 5-flanking region (131).


The PLG gene has a TATA-like sequence, TGTAA, at position -16
rather than a canonical TATA box. This sequence corresponds
exactly to the TATAA box of the homologous apo(a) gene, located
31 bp upstream from its transcription start site. Accordingly, the
TGTAA sequence is probably involved in the expression of the
TATA-less promoter in the PLG gene (131). Two main regions
conferring liver specificity of expression to the PLG gene were
found in the cis-acting element, AAAAATA, at -2194 to -2188 and
+48 to +61 (132). Several other putative regulatory transcription
elements are present in the 5-flanking region (reviewed in 131).
On the 3-flanking region, the location of the consensus
polyadenylation sequence AATAAA and of potential CAYTG signals
suggests the presence of three polyadenylation sites (13).

PLASMINOGEN ACTIVATORS
There are two physiologic PAs, tPA (tissue-type) and uPA
(urinary-type, UK), named after the original sources of purified
proteins. A further activation system, which is dependent on factor
XII, prekallikrein, and HMW kininogen, is known as the contact PA
pathway. There are also PAs from bacteria and the vampire bat,
which are included here because of their therapeutic use in
thrombolysis; they are also of mechanistic interest. The two human
proteases, tPA and uPA, are also implicated respectively in
neurobiology (133) and tumor biology (134). These aspects are not
considered further in this chapter, which focuses on PAs in
hemostasis.

TISSUE-TYPE PLASMINOGEN ACTIVATOR


tPA, a serine protease of 68 kDa (EC 3.4.21.68), also known as
vascular PA or extrinsic activator, is a glycoprotein of 527
residues. It exerts its effect primarily in the vascular system,
because it is produced and secreted by endothelial cells (135);
many other cells in culture also synthesize tPA. In normal plasma,
the antigen concentration of tPA is approximately 5 g perL, which
corresponds to approximately 70 pM (136). Most of the tPA is
present in a complex with its primary inhibitor, PAI-1 (137,138).

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

25 de 138

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

Structural Organization
The gene for tPA comprises 32.7 kb and is located on bands
p12-p11 on chromosome 8 (139), containing 14 exons (140) (Table
18-1). The mature protein exists in two forms of different length.
The N-terminus of Bowes melanoma cell tPA can be either glycine
or serine (141), but serine is nearly always the N-terminus of
recombinant tPA (142). In this chapter, the numbering is based on
the Ser-terminus of recombinant tPA because of the extensive
literature that exists on structure-function in this form of tPA. In
this numbering, glycine -3 probably represents the real N-terminus
of tPA.
The structure of tPA (Fig. 18-1) shows a finger and EGF domain,
and two kringle domains in the A chain, whereas the protease
domain is the B chain. The finger domain extends from residues 6
to 43 and is 34% identical to the first finger of bovine fibronectin.
It is involved in the binding of tPA to fibrin, as shown by mutants
lacking kringle domains, which still bound to fibrin (143).
Consistent with this, a degraded form of tPA that had lost the
N-terminal 12 kDa bound less well to fibrin than wild-type tPA
(144). The binding of the finger domain is independent of LBS, in
that it cannot be blocked by ACA (145). Structurally, the finger
domain is very similar to the seventh type 1 repeat of human
fibronectin (146).
The epidermal growth factor domain (residues 4492) is
structurally similar to other EGF structures, as shown by nuclear
magnetic resonance (NMR) (147). There is some evidence from
deletion mutagenesis that the EGF domain binds to the mannose
receptor and is involved in tPA clearance. Kringle 2 has affinity for
lysine, -amino acids such as ACA, and fibrin, whereas no
function has yet been ascribed to kringle 1 (143). The affinity for
the binding of ACA (a model for C-terminal lysine residues) and of
N-acetyllysine methyl ester (a model for intrachain lysine residues)
is approximately equal, suggesting that tPA does not prefer
C-terminal lysine residues (as PLG does) for binding. Intact tPA
and a variant consisting only of kringle 2 and the protease domains
were found to bind to the cyanogen bromide (CNBr) fibrinogen
fragment FCB-2, which also binds PLG and acts as a stimulator of

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

26 de 138

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

tPA-catalyzed PLG activation. In both cases, binding was


completely inhibited by ACA, pointing to the involvement of LBS
in this interaction (143).
The structurefunction relations among kringle 2 and -amino
acids and lysine have been studied in detail, using
P.343
NMR (148), microcalorimetry (149), crystallography (150), and
site-directed mutagenesis (42,149,151). The crystal structure of
kringle 2 resembles that of PLG kringle 4; there are, however,
differences in the lysine-binding pocket. The core of kringle 2 is
formed by a hydrophobic cluster of three tryptophan residues in
positions 25, 63, and 74, surrounded by aromatic and hydrophobic
side chains that form, at the surface of the kringle, a hydrophobic
grove. Ligand binding appears to rely mostly on the integrity of
Trp63 and Trp74, and an aromatic residue at position 76, which is
normally Tyr (152). Mutation of the critical amino acids Lys33,
Asp55, Asp57, or Trp72 markedly diminished binding to
lysineSepharose, or interaction with ACA, or both (153).
The catalytic domain of tPA is typical of other serine proteases,
with the catalytic triad His322, Asp371, and Ser478. The 2.3-
crystal structure of the protease domain revealed strong structural
similarity with other trypsinlike serine proteases, thrombin in
particular (154). The active site cleft is shaped and narrowed by
four surface loops. The loop around Arg299 exhibits five additional
residues, Arg298-Arg-Ser-Pro-Gly302, compared with
chymotrypsin. It projects out of the molecular surface as a
-hairpin and is of fundamental importance for the interaction with
PAI-1 (3). The 60-loop around Arg327 is similar to but shorter than
the corresponding loop in thrombin. Further loops are found around
Ser381 and Gly465. The fully solvent-exposed hydrophobic region,
comprising amino acids 420 to 423 of tPA, which forms a surface
loop near one edge of the active site of tPA, constitutes an
important secondary site for the interaction of tPA with PLG in the
absence of fibrin.
tPA is secreted as a single-chain molecule (sctPA) but can easily be
converted to the two-chain form (tctPA) by plasmin, which cleaves
the Arg275Ile276 peptide bond. Surprisingly, the single-chain

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

27 de 138

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

form is not a zymogen but a protease which, in the presence of


fibrin, is nearly as active as the two-chain form (155). The
nonzymogenicity of tPA has been explored in a series of
comparisons, with confirmatory mutagenesis studies (3,156). It
emerges that other zymogens have a triad Asp194, His40 and
Ser32 (chymotrypsin numbering), the His and Ser of which are
replaced in tPA by Phe305 and Ala292, respectively. The
constructed double mutant Phe305His and Ala292Ser was more
zymogenic (156).

Enzymatic Properties of Tissue-Type


Plasminogen Activator
tPA is unusually specific; its major substrate is PLG, in which it
cleaves the Arg561Val562 bond. It is an inefficient activator of
PLG in the absence of fibrin, but in its presence the activation of
PLG is greatly potentiated (157). Since Thorsen et al.'s original
observation that fibrin binds tPA but not urokinase (158), this
phenomenon has been much studied. Binding of sctPA and tPA is
roughly comparable, although the single-chain form may bind
slightly better (159). The K d for binding of tPA to fibrin clots in the
absence of PLG ranges from 140 nM to 400 nM (155,159,160). In
the presence of PLG, the affinity of tPA to fibrin increases
approximately 20-fold (K d 20 nM) (161). These observations are
explained by formation of a ternary complex comprising tPA, PLG,
and fibrin or binding of tPA to PLG, which, on binding to fibrin, has
taken on an open conformation. The isolated A chain of tPA was
indeed shown to bind to Glu- and mini-PLG with a K d of 100 nM
(162).
The kinetic parameters reported for the activation of PLG by tPA
show great variation. This is due to many potential experimental
variables, different tPA and PLG preparations, different substrate
concentrations, and, for studies using fibrin or fibrin derivatives,
the nature of the fibrin stimulator used. In the absence of fibrin,
K M values for the activation of GluPLG by tPA range from M to
slightly more than 100 M (155,163,164). In general, K M is three
to four times lower with tPA than with sctPA when the activation of
GluPLG is investigated in the absence of fibrin. This difference
disappears when fibrin is present and K M values typically are two

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

28 de 138

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

orders of magnitude smaller, with only moderate increases of K c at .


Values for K M in the presence of fibrin range from 0.16M to 1.1
M PLG, and k c at values from 0.1 per second to 1.1 per second
(155,164). Several authors found nonlinear enzyme kinetics
(162,163,165), and there appear to be two phases in the activation
of GluPLG by tPA in the presence of fibrin, with an initial K M of
1.05 M and k c at 0.15 per second. Later in the process, as new
high-affinity binding sites for PLG and tPA are exposed in partially
digested fibrin (61,62,63,166,167), K M decreased to 0.07 M
whereas the k c at was unchanged (165). The important message
from these studies is that PLG is not activated by tPA except in the
presence of fibrin, because it is only in the presence of fibrin that
the K M for the reaction is consistent with the circulating
concentration of 2 M.
Distinct sites in the fibrin molecule have been identified as
accelerating PLG activation; notably they are not available for
binding in fibrinogen but become exposed in fibrin (168). The
D-region of fibrin binds both PLG and tPA in a region that
encompasses the sequence A148-160, and tPA alone, in a site
including 312-324. Both these associations are of low affinity and
in the circulation the A148-160 site would bind PLG, which is
present in much higher concentrations than tPA. Another binding in
the D-region encompasses 311-336 and 337-379, which are
linked by a disulfide bond (169). Antibodies have revealed that
there is a tPA binding site that includes 312-324 (170). Higher
affinity sites for binding tPA and PLG are in the C-terminus of the
chain, within the region A392-610 (171). The conformational
changes that occur on fibrinogen cleavage and fibrin assembly
(172) reveal new sites for binding of tPA and PLG, and for
enhancement of PLG activation.
Huge numbers of mutations have been engineered into tPA to
change its characteristics, with the aim of making it an even more
effective therapeutic agent. The most striking of these is the series
of mutations to make tPA resistant to PAI-1. Madison et al.
identified the importance of interactions between the positively
charged tPA sequence 298302 and the negatively charged PAI-1
sequence 350355 (173). Mutagenesis of Arg298, 299, and 304
resulted in a mutant that was inhibited 120,000 times less rapidly

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

29 de 138

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

by PAI-1 than wild-type tPA (174). These observations led, in part,


to the development of the new tPA mutant tenecteplase (TNKtPA),
in which the sequence 296299 (Lys-His-Arg-Arg) is replaced by
four alanines (175). This variant also exhibits slower clearance, by
virtue of its changed glycosylation, and its enhancement of PLG
activation is more selective for fibrin above fibrinogen and the
fibrin fragment DDE (176).

Tissue-Type Plasminogen Activator


Receptors and Binding Proteins
A number of structurally unrelated components that bind tPA have
been described. Some have the potential to localize tPA on a
surface to which PLG also binds, and therefore to potentiate
activation. Many of the proteins in this group contain a C-terminal
lysyl group and therefore also are receptors for PLG, as already
discussed (69). This group includes 42-kDa annexin II (69,177),
45-kDa actin (178), heparan sulfate and chondroitin sulfatelike
proteoglycans (179), cytokeratin 8 and 18 (180), and tubulin
(181). Overexpression of the tPA receptor annexin II, as it occurs
in patients with promyelocytic leukemia exhibiting the t(15;17)
translocation in the
P.344
leukemic cells, is associated with a hyperfibrinolytic state. The
abnormally high levels of annexin II bind PLG and tPA and
generate plasmin, with consumption of 2 -antiplasmin, leading to
bleeding (182). Interestingly, other studies show that such
patients also have abnormally high fibrinolytic activity that is
clearly due to uPA (183).
Several as yet poorly defined tPA receptors have been identified on
endothelial cells, and it is not clear yet whether some of these are
identical with each other (184). In addition to actin, human
umbilical vein endothelial cells (HUVEC) express two further
37-kDa and 45-kDa tPAbinding proteins (178), whereas another
study found a 20-kDa tPA binding protein, which was characterized
and purified (185). This protein did not interact with PLG, but
binding was inhibited by -amino acids, suggesting that the
lysine-binding site of tPA is involved, but the purified protein was
not recognized by antibodies to annexin II or -enolase. In the

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

30 de 138

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

presence of this receptor, PLG activation by tPA was enhanced


90-fold. A receptor that interacts with the B chain of tPA and
potentiates PLG activation approximately 100-fold has been
described on vascular smooth muscle cells (186). It has now been
defined as type-II transmembrane protein p63 (CKAP4) (187).

Glycosylation of Tissue-Type Plasminogen


Activator
tPA has four potential glycosylation sites, of which three are
occupied in type I (Asn117, Asn184, and Asn448) and two in type
II tPA (Asn117 and Asn448), making it 3 kDa smaller (163).
Residue 117 is predominantly N-glycosylated with
oligomannose-type structures, whatever the source. Residues 184
and 448 are predominantly associated with complex-type
structures in recombinant tPA and in tPA isolated from fibroblast
cell lines, but with both complex- and oligomannose-type
structures when isolated from melanoma cells (188). Glycosylation
at residue 184 influences the biological properties of tPA. The type
II form has a higher affinity for lysine (and fibrin) and a higher
fibrinolytic activity (189). tPA, like uPA, has an O-linked fucose on
Thr61 in the EGF domain (190) and this affects binding to HepG2
cells and may be relevant to clearance (191).

Polymorphisms
The best-studied polymorphism is an Alu insertion/deletion
polymorphism. Most clinical studies including a large prospective
study (192), found no correlation with acute myocardial infarction
or stroke, but one casecontrol study found that homozygotes for
the insertion had twice as many cases of acute myocardial
infarction as homozygotes for the deletion (193). The homozygous
insertion polymorphism is also associated with a high rate of
release of tPA in response to stress (194). Further study of the
promoter identified additional polymorphisms, one of which
(-7351C> T) has been shown to be functional at the level of
transcription, with decreased release of tPA in individuals carrying
the T allele (195).

Synthesis, Release, and Clearance of

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

31 de 138

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

Tissue-Type Plasminogen Activator


tPA is produced by many different cell types in culture, and its
synthesis can be upregulated by diverse agents. Endothelial cells
are a principal source and, in these cells, release rather than
synthesis is also important in controlling local tPA. Finally, tPA is
very rapidly cleared from the circulation. These three aspects are
now reviewed briefly.

Synthesis
tPA expression in cultured cells can be stimulated through multiple
intracellular signaling pathways. Vasoactive substances, such as
thrombin and histamine, increase tPA synthesis in HUVEC, acting
through their G-coupled receptors and the protein kinase C (PKC)
pathway (196,197,198). Steroid hormones (199,200) and retinoids
can increase synthesis of tPA (201,202). It is important to note
that not all endothelial cells synthesize tPA in vivo and that
expression is restricted to small vessels (203,204,205).
The tPA gene contains common transcription elements, with three
TATA boxes; a CAAT box is present at position -112 to -116. DNase
I protection analysis of the tPA gene promoter in human
endothelial- and phorbol-stimulated HeLa cells revealed several
protected regions. A cytotoxic factor/nephritic factor-1like
element is at -92 to -77; three Sp1 binding sites are at positions
-72 to -66, -48 to -39, and +60 to +74; a cyclic adenosine
monophosphateresponsive element CRE-like/tetradecanoyl
phorbolacetateresponsive element TRE-like element is between
-102 and -115; and three GC/GT boxes are between -43 and +68
(196,206,207). Several of these cis-acting elements have been
shown to be involved in constitutive and phorbol esterinduced
expression of tPA mRNA. Between -2,288 to -2,129 and -2,390 to
-2,289 are two further regulatory elements necessary for
constitutive expression of the tPA gene in Bowes melanoma cells
(208). Far upstream at -7.3 kb is located a functional retinoic acid
response element consisting of a direct repeat of the GGGTCA
motif spaced by five nucleotides (209). The region -7,145 to
-9,758 comprises a multihormone-responsive enhancer that is
activated by glucocorticoids, progesterone, androgens,
mineralocorticoids, and 1,25-dihydroxyvitamin D 3 but not by

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

32 de 138

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

estrogens (199,200). A nuclear factor NF1 site, 600 bp upstream


from the transcription start site, acts as a repressor of tPA
expression and confers cell specificity to tPA expression (205).

Release
Acute release of tPA is established from early studies on venous
occlusion for 10 or 20 minutes, a procedure widely used in patients
(210,211,212). During venous stasis, at midway between systolic
and diastolic blood pressure, venous outflow from the occluded
segment decreases more than the (diminished) arterial inflow and
represents approximately one third of remaining arterial inflow
(210). Much of the locally produced tPA therefore remains in the
occluded limb and is not cleared. Young patients with a history of
recurrent venous thromboembolism exhibited an abnormal
response to venous occlusion and this could be attributed either to
subnormal tPA release or, more often, to elevated PAI-1
(137,211,212). In isolated cases with severe forms of von
Willebrand disease, a complete lack of response to venous stasis
and the infusion of 1-deamino-8-D-arginine vasopressin (DDAVP)
was found (213). These patients may have a functional or
structural defect in their endothelial cells. It should be noted,
however, that release of von Willebrand factor (VWF) and tPA
occur by different mechanisms (214). Lower PA activity is found in
leg veins than in proximal veins. It has been estimated that the
synthesis and release rate of tPA during forearm occlusion ranges
between 0.5 to 1.1 ng per mL of blood, but only 0.08 to 0.11 ng
per mL in an occluded leg (210). Increased hydrostatic pressure in
the leg veins may be responsible for the diminished production of
tPA. Applying a 20-minute venous occlusion test, patients who
were immobilized were found to release considerably more tPA
from leg veins after 12 to 33 days of recumbency than at the
beginning of immobilization (215). The low content of tPA in
human calf veins therefore may be one etiologic factor for
development of deep venous thrombosis.
P.345
There is evidence that continuous release of tPA from the
vasculature can diminish a thrombotic response. The human tPA
gene was cloned into an adenoviral vector under the control of the

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

33 de 138

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

Rous sarcoma virus (RSV) promoter and the construct was tested
in an in vivo model of arterial thrombosis. All animals developed
obstructive thrombosis except those treated with the viral vector;
this was actually more effective than tPA therapeutic regime (216).
This echoes two independent studies in the 1980s of patients who
had consistently raised tPA activity as the cause of a hereditary
lifelong hemorrhagic disorder (217,218). One of these (217) had a
total absence of fibrin deposits or thrombi, despite widespread
arterial disease; he died of cerebral bleeding. The underlying cause
was not a deficiency of PAI-1 or any other known inhibitor. No
further cases have been reported, despite an intensive search in
many laboratories.
The mechanism of release has been studied in cultured endothelial
cells, from which much of the tPA is released continuously, but
there is an additional regulated pathway, where tPA is stored in
specialized storage vesicles (219,220,221,222). Extracellular
stimuli then trigger the release of stored tPA to the cell surface.
The amount of tPA stored in storage vesicles in various cells, such
as endothelial, neuroendocrine, and adrenal chromaffin cells, is
large (220,222). In rats, the tissue stores of tPA were calculated
to be sufficient to maintain a steady-state plasma level of tPA for 2
days in the absence of protein synthesis (223). Compounds that
triggered release within a few minutes include bradykinin,
histamine, eledoisin, acetylcholine, -adrenergic agents,
platelet-activating factor, endothelin, calcium ionophore A-23187,
and acidosis (224,225). These compounds induce calcium influx
into the endothelial cell and activate G-proteincoupled receptors
(226). Some interventions that cause elevated tPA act through
decreased clearance, as discussed in the next section. Among
these are exercise and -adrenergic agents (227). In all situations
in which epinephrine levels are increased, such as stress, anxiety,
and exercise, tPA antigen levels increase (138,228).
DDAVP was widely used in the 1980s to study the release potential
for tPA in patients with idiopathic or recurrent thrombosis
(215,229,230,231). Earlier observations had suggested that DDAVP
acts through a release of a pituitary PAreleasing hormone (229).
Comparison of the effects of DDAVP and sodium nitroprusside
showed that nitroprusside produced an even greater increase of

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

34 de 138

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

forearm blood flow but no increase of tPA, but that DDAVP indeed
stimulated tPA release from the vascular bed (232). The increase
of circulating tPA levels after intraarterial infusion of acetylcholine
and methacholine was shown to be mediated by muscarinic
receptors (232,233,234). Bradykinin and substance P both induced
tPA release after being infused into the forearm of volunteers
(235,236), and the releasable pool of tPA, rather than the
availability of PAI-1 to inhibit it, was the key factor (237).

Clearance
Free tPA and tPA in complex with inhibitors are rapidly removed
from the circulation and bound to receptors on endothelial cells
and hepatocytes. In normal individuals, the t 1/2 of tPA is
approximately 3 minutes (238,239). Studies on isolated rat
hepatocytes suggested that the complex was cleared more rapidly
than free tPA (240) but, in vivo in humans, clearance of tPA was
calculated to be faster in subjects with low PAI-1 (3.5 minutes)
versus high PAI-1 (5.3 minutes; P = 0.006), consistent with t 1/ 2 of
2.4 minutes for free tPA and t 1/ 2 of 5.0 minutes for tPA/PAI-1
complexes (P = 0.006) (241). Because of its clearance by the liver
and the dependence of clearance on liver blood flow
(242,243,244), the t 1/ 2 may be considerably prolonged in patients
with hepatic cirrhosis (239,245).
Early studies showed that the uptake of tPA by liver endothelial
and by Kupffer cells was inhibited by ovalbumin, leading to the
identification of the mannose receptor as a major tPA receptor
(246). Binding of ligands to the mannose receptor is pH sensitive
and dependent on Ca 2+ ions. Binding is inhibited by
mannose-albumin, mannan, D-mannose, and L-fucose, and
particularly by cluster mannosides of the composition M 6 L 5
(246,247,248). tPA has a high affinity (K d 1 to 4 nM) for the
mannose receptor compared to other glycoproteins, the K d for
most being 60 to 600 nM (248). In mice, the administration of
estradiol leads to increased expression of the mannose receptor
and increased clearance of tPA (249). Mutants of tPA have been
engineered for slower clearance by replacement of Asn117, which
is normally glycosylated, with Gln. Such mutants, including
TNKtPA, are cleared more slowly than wild-type tPA. It is not

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

35 de 138

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

known whether the fucose residue at Thr61 in the EGF domain also
binds to the mannose receptor or to the previously described
fucose receptor (250).
The other major pathway for tPA clearance involves the
LRP/ 2 -macroglobulin receptor (251,252). The 600-kDa LRP, the
largest plasma membrane protein described, comprises 4,525
residues. It mediates the clearance of apolipoprotein Eenriched
chylomicron remnants, toxins, cytokines, complexes of
2 -macroglobulin with proteases from all subclasses, and free tPA
and tPA/PAI-1 and uPA/PAI-1 complexes. Complex formation of tPA
with PAI-1 increases the rate of clearance by LRP by at least one
order of magnitude compared with that of free tPA (253). The
binding sites for tPA/PAI-1 and for uPA/PAI-1 complexes are
situated in the second complement-type domain cluster of LRP
(254,255,256). The receptor-associated protein (RAP) inhibits
endocytosis of most ligands to LRP (251,254,257).
Other LRP-like multiligand receptors, such as glycoprotein 330 and
the 130-kDa very low density lipoprotein (VLDL) receptor, also can
mediate the clearance of free and PAI-1 complexed tPA
(258,259,260). For efficient uptake and clearance of several
ligands, LRP works in cooperative fashion with coreceptors, for
instance uPAR for the LRP-mediated degradation of uPA/PAI-1
complexes (261,262). Some monocytelike cell lines that expressed
LRP on the cell surface were unable to degrade tPA, leading to the
idea that a coreceptor was necessary for efficient degradation of
tPA by LRP (263).

Deficiency
Congenital deficiencies of tPA or uPA have not been described in
humans and were assumed to constitute a lethal condition, until
these genes were successfully deleted in mice (264).
Unexpectedly, mice with a disrupted tPA gene developed normally.
Microscopic examination revealed mild glomerulonephritis. In a
plasma clot lysis assay, fibrinolytic activity was lower than in
wild-type animals. After endotoxin-induced thrombogenic
stimulation, tPA -/ - mice had more extended thrombotic lesions than
tPA + /+ mice. The abnormalities found were rather mild, but mice

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

36 de 138

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

with a double deficiency of tPA and uPA had severe spontaneous


thrombosis and exhibited impaired health, reduced survival, and
diminished fertility (264,265). It can be concluded that, in mice,
there is some redundancy of biological functions with respect to
PAs. On the other hand, it is obvious that a complete knockout of
the fibrinolytic activators leads to severe disease. It is not known
to what extent these results apply to human pathophysiology. The
double knockout has a similar phenotype to the PLG knockout,
consistent with tPA and uPA being the major mammalian activators
of PLG. These mice have had a profound impact on understanding
other functions of tPA, in areas as diverse as reproduction (266),
neuronal development (133,267) and learning (133,268).
P.346

URINARY-TYPE PLASMINOGEN ACTIVATOR


The urinary-type PA (EC 3.4.2.73; urokinase, uPA) is found in
urine at 40 to 80 g per L (269) and is synthesized by several cell
types, particularly cells with a fibroblastlike morphology, but also
by epithelial cells (270) and monocytes and macrophages
(271,272). Stimulation by endotoxin or tumor necrosis factor
causes cultured endothelial cells to produce uPA, and synthesis has
also been demonstrated by the human endothelium in vivo
(273,274). uPA activates PLG by the cleavage of Arg561Val562,
as does tPA, but importantly can do so in the absence of fibrin.
This characteristic has generated the view that tPA functions in
fibrinolysis in the vasculature, whereas uPA's primary role is in
processes such as degradation of extracellular matrix and cell
migration, with consequences for wound healing, inflammation,
embryogenesis, and invasion of tumor cells and metastasis
(275,276). These clear functions of the uPA system do not argue
against its importance in fibrin degradation, functions that have
emerged from both human and animal studies, as discussed in
subsequent text.

Structural Organization
The complete primary amino acid sequence of uPA was determined
in 1982 (277). The cDNA and the genomic DNA of uPA have been
isolated and their nucleotide sequences established (278,279). The

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

37 de 138

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

human uPA gene is 6.4 kb and on chromosome 10q24 (280). It


contains 11 exons, and the intronexon organization of the uPA
gene closely resembles that of the tPA gene. uPA is produced as a
54-kDa, single-chain glycoprotein, 411 residues long and
containing three domains: an epidermal growth factor domain, a
kringle, and a protease domain (Fig. 18-1). The EGF domain
interacts with the uPAR found on many cells, as discussed in
subsequent text. The kringle domain has no affinity for fibrin, and
its function has been unclear; a recent report shows that it
stabilizes the interaction of uPA with uPAR (281). The protease
domain has the residues His204, Asp255, and Ser356 as the
catalytic triad, and sequence identity with tPA is approximately
40% (140). uPA is normally glycosylated at Asn302, whereas a
covalently attached single monosaccharide, fucose to Thr18 (282),
appears to affect the mitogenicity of uPA (283). There are two
phosphorylation sites on Ser138 and Ser303. The phosphorylated
form diminishes uPA's interaction with cells and PAI-1 (284).
Activation of single-chain urinary plasminogen activator (scuPA,
also called proUK) to the active enzyme, uPA (UK), is by cleavage
of Lys158Ile159 (285). Many enzymes can cleave this bond but
the most important to hemostasis are plasmin (286), factor XIIa,
and kallikrein (287). The activation of scuPA in PLG-deficient mice
was characterized as being due to kallikrein (288). Cleavage close
to this site results in inactive uPA. Thrombin cleaves
Arg156Phe157 in scuPA (287,289,290) and the inactive product
can be activated by release of the N-terminal dipeptide of its B
chain by cathepsin C or, albeit slowly, by plasmin (285,291).
Granulocyte elastase and cathepsin G cleave the Ile159Ile160
peptide bond; again, the resulting two-chain urinary-type
plasminogen activator (tcuPA) is inactive (292).
uPA was originally characterized as having two active forms:
HMW-UK and LMW-UK (293). These are, respectively, full-length
active uPA and a processed form, cleaved at Lys135Lys136 to
yield the amino-terminal fragment (ATF), which interacts with
uPAR, leaving 21 residues of the A chain and an intact B chain.
Cleavage within the uPAR binding region (e.g., of the Lys23Tyr24
bond) abolishes uPA receptor binding and therefore modulates cell
surface uPA activity (294,295).

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

38 de 138

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

The conformation of uPA and of its domains has been studied by


NMR (296,297). Essentially, the domains are similar to those in
other serine proteases (Fig. 18-1) with few considerable
differences. The crystal structure of the catalytic domain of
recombinant, nonglycosylated human uPA has been solved at a
resolution of 2.5 (298). At six positions, insertions of extra
residues in loop regions create unique surface areas. The
interaction of Lys300 with Asp335 in the flexible loop region
297313 stabilizes the conformation of scuPA (299). The positively
charged residues 179184 in a surface loop interact with PAI-1
(297). A number of modified uPA molecules have been crystallized
and used as a basis for drug design (300,301).

Enzymatic Properties of Urinary-Type


Plasminogen Activator and Single-Chain
Urinary Plasminogen Activator
The question of whether scuPA is a true zymogen was hotly
debated for several years, a controversy that was fueled by the
problem of trace plasmin or uPA contamination, with reciprocal
activation of scuPA and PLG (302). The issue has been resolved by
studies on mutant forms of the proteins and by examining scuPA in
the absence of PLG. For instance, the Lys158Glu mutant of scuPA,
which cannot be cleaved by plasmin into uPA, still converted PLG
to plasmin (303). The emerging consensus is that scuPA has
approximately 0.4% of the activity of uPA (304,305,306,307).
Therefore, scuPA lies on the spectrum between PLG, a true
zymogen, and tPA, which is almost fully active in its single-chain
form. Lys300 and Asp355 are key structural elements of scuPA's
enzymatic activity. Lys300, situated in the flexible loop region
297313, forms a weak interaction with Asp355 adjacent to the
active site Ser356 and pulls Ser356 close to the position found in a
fully active protease. Site-directed mutagenesis of Lys300 to Ala
resulted in a 40-fold lower activity than in wild-type scuPA (308).
Likewise, a mutation of Asp355 to Asn resulted in a 270-fold
reduction of scuPA activity (299). When the flexibility of the
297313 loop was enhanced, intrinsic catalytic activity was
reduced. By contrast, when the loop was made less flexible,
intrinsic catalytic activity increased (309).

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

39 de 138

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

Urinary-Type Plasminogen Activator


Receptor
A cellular receptor for uPA (uPAR; CD87) was first demonstrated on
freshly collected human monocytes (310) but is expressed on
several cell types. The gene for uPAR (gene code PLAUR) is on
chromosome 19 and consists of seven exons and six introns
extending over 23 kb (311,312). Constitutive expression of the
uPAR gene is dependent on the -141 to -61 region (313), and
expression can be increased by tumor promoters, growth factors,
cytokines, and hormones. Interestingly, it can be upregulated by
one of the plasmin reductases, phosphoglycerate kinase (75,314).
The 1.4-kb mRNA encodes a signal peptide of 22 and a protein of
313 residues, with a glycosylphosphatidyl inositol (GPI) anchor, by
which the receptor is attached to cell surface phospholipids (see
Fig. 18-6). The receptor consists of three homologous domains,
and high-affinity binding of uPA to the receptor is mediated by
region 1, but region 2 and 3 enhance this binding (315,316,317).
scuPA and uPA bind to uPAR receptor by the ATF (315). The
binding is of high affinity (K d 10 - 9 to 10 -11 depending on the cell
type) (318) and can localize and enhance uPAmediated PLG
activation on cell surfaces (71). The mechanism is a colocalization
of scuPA and PLG; uPAR does not play a catalytic role
P.347
(319). Binding can also elicit a number of intracellular signaling
responses, acting through the extracellular regulated kinase
(ERK)/mitogen activated protein kinase (MAPK) pathway (320).
Because uPAR is not a transmembrane protein, other components
must participate to transmit a signal to the intracellular
compartment. Several proteins are known to bind uPAR, including
vitronectin and integrins in complex with caveolin (321). Such
interactions regulate cell adhesion and are discussed further in the
context of PAI-1 in Chapter 19. On some cells, the complex
uPA/uPAR is associated with Endo 180, also known as
uPAR-associated protein (uPARAP) (322), and is involved in
collagen IV internalization (323). Receptor-bound uPA still is
susceptible to inactivation by PAI-1, but the inhibition is somewhat
less efficient than in free solution (319). The uPA/PAI-1

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

40 de 138

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

enzyme/inhibitor complexes are internalized, and the uPAR is


recycled to the surface. This process requires the cooperation of
LRP (262,324).

FIGURE 18-6. A simplified diagram of the domain structure of


urinary-type plasminogen activator receptor (uPAR) and its
attachment to the cell-surface glycosylphosphatidylinositol
(GPI) membrane anchor. The N-terminal portion of the uPAR is
composed of three domains. Potential glycosylation sites are
indicated by . The C-terminal amino acids are depicted by
squares. The linkage between the protein and the glycolipid is
through a phosphoethanolamine that forms an amide bond with
the a-carboxyl group of the protein and a phosphodiester bond
with the glycan portion of the glycolipid. (PI-PLC,

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

41 de 138

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

phosphatidylinositol-specific phospholipase C.) (From Ploug M,


Behrendt N, Lber D, et al. Protein structure and membrane
anchorage of the cellular receptor for urokinase-type
plasminogen activator. Semin Thromb Hemost
1991;17:183193, with permission.)

Urinary-Type Plasminogen Activator in the


Circulation
scuPA is found in plasma in concentrations of 2 to 4 ng per mL
(325,326), which seem to remain fairly stable. In contrast with tPA
and PAI-1, there is little circadian fluctuation in plasma uPA (327).
Venous occlusion did not increase uPA antigen once values were
corrected for the increase of hematocrit (136), but there are
reports of increases of uPA antigen after venous stasis (326),
DDAVP infusion (328), and strenuous physical exercise (329). The
activation of scuPA in plasma has not been studied in any detail,
and uPA activity is not normally detected in plasma, but both
leukocyte-associated and free scuPA are elevated in leukemia
(183) and other disorders, including liver disease (330). Plasma
uPA is cleared quickly, with a t 1/ 2 of approximately 7 minutes, in a
manner that depends on hepatic blood flow (331). The LRP system
binds and internalizes scuPA and uPA-PAI-1 complexes
(251,261,262). The asialoglycoprotein receptor, located exclusively
on parenchymal liver cells, also removes nonsialated uPA from the
circulation (331).
The activity of scuPA is increased by two orders of magnitude when
it is bound to uPAR on the surface of monocytes (332,333), which
is now understood to reflect colocalization with PLG rather than a
direct catalytic effect of uPAR (319). This local activity is relevant
to thrombus stability, into which monocytes migrate (334) and
express fibrinolytic activity (335). In freshly formed model
thrombi, there is considerable spontaneous generation of
fibrinolytic activity, most of which can be inhibited by antibodies to
uPA (336). Polymorphonuclear cells, primarily neutrophils,
generate this local uPA activity, apparently on uPAR (337); plasma
1 -antitrypsin is crucial in protecting the activity from neutrophil

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

42 de 138

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

elastase (338). The integrin M 2 is important in the generation of


such local activity (339). When scuPA is present on uPAR, for
instance on monocytes, it can associate with PAI-1 and other
serpins, not in a classic covalent complex but in a reversible
manner (340). This has given rise to the ideas that receptor-bound
scuPA initiates proteolytic activity
P.348
and that conversion to uPA is primarily a way of achieving
inhibition by PAI-1 and other inhibitors and thereby regulating the
activity (341). Platelets also contain uPAR (342) and a further
70-kDa protein, distinct from uPAR bound uPA in resting platelets
with a K d of 43 nM (343). On other cells, too, cell-associated PLG
activation by uPA is not mediated exclusively by uPAR and other
receptors may come into play (344).
Although it does not bind to fibrin, scuPA has some fibrin
specificity in that it degrades fibrin with less systemic degradation
of fibrinogen and other plasma proteins than does uPA. It therefore
is a more potent thrombolytic agent than uPA (306,345). The
explanations for this include the different but complementary
mechanisms by which tPA and scuPA induce fibrinolysis (346). tPA
activates GluPLG, which is bound to an internal lysine, probably
A Lys157, which is exposed in nondegraded fibrin. As soon as
some fibrin is digested by plasmin, new C-terminal lysyl residues
are generated, increasing the binding of GluPLG (61,63) to fibrin.
scuPA binds with high affinity to PLG (K d approximately 65 nM) and
appears to activate selectively GluPLG, which is bound to
C-terminal lysines in partially degraded fibrin (347,348). The
generation of small amounts of fibrin-bound plasmin also converts
scuPA to uPA, explaining the typical lag phase observed for
fibrinolysis by scuPA. Efficient scuPA-mediated lysis occurs after
some fibrin has been degraded by tPA. Therefore, sequential
administration of tPA followed by scuPA resulted in increased
thrombolysis in a rabbit jugular vein thrombosis model; the
reverse order was less effective (349).
During coagulation, there is activation of the contact system
generating kallikrein also. Because kallikrein is an efficient
activator of scuPA (287,350), it is likely that scuPA is activated
wherever activation of the contact phase of coagulation has

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

43 de 138

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

occurred. This process also takes place in the absence of PLG. In a


plasma milieu, clot lysis by scuPA is enhanced approximately
20-fold after dextran sulphate mediated activation of the contact
phase (350). It is not entirely clear if this in vitro generated
activity of kallikrein fully explains the contact system of PLG
activation (351). Part of the dextran sulfate mediated activation
may also be due to factor XII activity. Factor XIIa is a poor
activator of PLG but is present at high concentrations in plasma
(352).

Regulation of Expression
The expression of scuPA has been reviewed in detail (353). The
gene is regulated by several elements, with a typical TATA box 25
bp upstream of the transcription initiation site. A region of 500 bp
contains potential binding sites for the transcription factor Sp1.
Two putative AP-2 binding sites close to the transcription initiation
site are responsible for protein kinase A dependent induction of
uPA expression. Two NF-kB elements occur at -1,580 and -1,865
bp. The latter site acts as a repressor, as do two further negative
regulatory sites, one situated between -1,824 and -1,572 bp, the
other involving an enhancer-dependent cell-specific silencer region
between -660 and -536 bp. The promoter is strongly enhanced by
a region approximately 2,000 bp upstream of the transcription
initiation site. This enhancer consists of an upstream PEA/AP-1A
site and a downstream AP-1B site. All three sites are required for
induction of uPA gene transcription by a variety of extracellular
stimuli, such as phorbol ester, EGF, okadaic acid, and cytoskeleton
disruption (354). Synergism between PEA/AP-1 and AP-1B depends
on the integrity of an intervening cooperation mediator site that
contains several sequences for binding of urokinase enhancer
factors (355,356). In the 3 untranslated region UTR stretching
from bp 1,368 to 2,260 of the human uPA gene, fragments were
identified that exerted a positive or negative influence on chimeric
reporter genes. Fragments 1,999 to 2,190 behaved like
transcriptional enhancers, whereas fragments 1,532 to 1,723 had
an inhibitory effect on the expression of uPA (357). At least three
elements were found in the 5 UTR that confer instability to the
uPA mRNA (358).

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

44 de 138

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

Urinary-Type Plasminogen
Activator/Urinary-Type Plasminogen
Activator Receptor Deficiency
There are no known deficiencies of uPA or uPAR in humans, but
mice with these disrupted genes reveal a number of interesting
insights. uPA -/ - mice, like tPA - / - mice, developed some fibrin
deposits but did not spontaneously clot; challenge with endotoxin
led to more venous thrombosis than in uPA + / + mice (264,265).
Lysis of plasma clots was normal (264), but monocyte-dependent
lysis was greatly decreased by the absence of uPA (359). This is
consistent with the earlier findings that fibrin resolution was
affected in mice deficient in uPA but not by deficiency in tPA or
uPAR (360). Much lower plasmin generation, and hence matrix
metalloproteinase activation, was observed in mice with combined
apoE and uPA deficiency, so that they were protected from the
atherosclerotic lesions that were induced in apoE - /- mice (361).
Wound healing in uPA - /- mice is greatly disturbed and, after
injury-induced depletion of vascular smooth muscle cells,
repopulation of the injured area with smooth muscle cells was slow
in uPA - / - mice (362). Perhaps surprisingly, the healing process
does not involve binding to uPAR (363), and there are several
other reports on uPAR-independent but uPA-dependent phenomena
(364) as well as some in which both uPA and uPAR deficiency
produce related effects (365). Some aspects of uPAR biology, such
as the delayed clearance of platelets in uPAR - /- mice (366),
macrophage infiltration into the vessel wall (367), and cancer cell
growth and invasion (368), are independent of uPA. Studies on
mice are helping to unravel which effects of uPA and uPAR operate
independently of each other and which effects require their
common functions.

OTHER PLASMINOGEN ACTIVATORS


Streptokinase
In 1933, it was shown that hemolytic streptococci produced an
extracellular protein that could induce lysis of human plasma clots
(369), and it has been a useful thrombolytic drug over several

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

45 de 138

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

decades. Streptokinase and other bacterial activators have


presumably evolved for the purpose of invasion of the mammalian
host (370). SK is not an enzyme; rather, it acts by forming a 1:1
complex with PLG, which undergoes a change of conformation to
expose an active site in the altered PLG molecule (371). The
modified PLG in the equimolar complex (SK-PLG') autocatalytically
converts to plasmin, and the activated complex is a potent PA
(372). It retains the fibrin binding of the PLG moiety with
consequent relative protection from inhibition by 2 -antiplasmin
(373). The C-terminus of SK regulates the PLG binding and
activation, and mutational studies have revealed the essential
features (374).
SK is a 47-kDa protein, containing no cysteines. NMR and
crystallographic studies revealed a three-domain structure
(375,376). The binding site for a kringle in PLG is located at the
tip of a fully exposed hairpin loop in the -domain of SK
(377,378). The physicochemical properties of SK and the methods
for its assay are reviewed (379); a new reference method that
allows direct comparison with other PA has been developed (380).
P.349
SK is highly antigenic. Most apparently healthy persons have
anti-SK antibodies in their blood, owing presumably to previous
infections with -hemolytic streptococci, and anti-SK titers may
rise several hundredfold after administration to humans (381,382).
After intravenous injection of SK into humans, dogs, or mice, SK is
cleared from the circulation with a t 1/2 of 15 to 30 minutes (383).
Clearance mostly takes place in the liver, primarily by the liver
2 -macroglobulin receptor (383), now known to be identical to LRP
(251).

Anisoylated PlasminogenStreptokinase
Activator Complex
Plasmin generated in whole blood or plasma is immediately
inactivated, but plasmin bound via its high-affinity LBS to fibrin is
inactivated at a much lower rate by 2 -antiplasmin. In the search
for more efficient thrombolytic agents, researchers acylated the
active site center of plasmin and of the SKPlg' activator complex,

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

46 de 138

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

in which a conformational change also exposes the active


plasmin(ogen) site. Coupling of p-anisoyl to the active site serine
of the PLG portion of SKPlg' led to the development of the
anisoylated PLGSK activator complex (APSAC, Eminase). Its
catalytic site is blocked, but its affinity to fibrin via the LBS is
preserved. APSAC deacylates with a t 1/ 2 of approximately 40
minutes at 37C (384). As it binds to fibrin, it is feasible to
administer this drug in the form of a bolus injection. Evaluation in
experimental animal models showed that APSAC had better clot
specificity than SK, in that it caused less fibrinogenolysis and
consumption of PLG and 2 -antiplasmin and had a lesser
hypotensive effect in vivo (385). Clinical studies showed that
APSAC is an effective thrombolytic agent, but the systemic effects
are comparable with SK.

Staphylokinase
Staphylokinase (SAK) a protein with an M r of 15,000, produced by
Staphylococcus aureus, has been known for more than 40 years to
have thrombolytic properties (386,387). As recombinant
techniques made production possible in larger quantities, there
was renewed interest in its thrombolytic efficacy (388). Like SK,
SAK is not an enzyme but forms a stable stoichiometric complex
with PLG (372). In the absence of fibrin, it does not activate PLG.
Trace amounts of plasmin, such as may be found on the surface of
a thrombus, convert the SAKPLG complex to SAKplasmin
complex that is a highly fibrin-specific PA. The complex binds to
fibrin via the LBS of PLG, and the bound activator complex is
somewhat protected from inactivation by 2 -antiplasmin. The
molecule is quite antigenic, and extensive mutagenesis has been
undertaken to decrease this property (389). Coupling of SAK
mutants to polyethylene glycol of M r 20,000 was shown to increase
the t 1/ 2 of the complex severalfold, which would permit single
bolus thrombolytic therapy in humans (390). Several clinical
studies with SAK have been reported. In peripheral arterial
occlusions and in patients with myocardial infarction, SAK
compared favorably with tPA (391,392).

Vampire Bat Plasminogen Activator

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

47 de 138

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

Many hematophagous animals, such as leeches, hookworms,


mosquitoes, and the vampire bat, produce substances that
interfere with the clotting or fibrinolysis mechanisms. In the
1960s, a potent PA was found in the saliva of the common vampire
Desmodus rotundus (393). The bat lives exclusively on blood and
uses the enzyme in its saliva to keep blood in a fluid state. The D.
rotundus salivary plasminogen activator (DSPA) was cloned
independently by two research groups and shows more than 70%
identity with human tPA (394,395). Three major forms were
identified: DSPA-, which contains a finger, EGF, kringle, and
protease domain; DSPA-, which lacks the finger domain; and
DSPA- which has just a kringle and a protease domain.
DSPA- comprises 441 residues and, like SAK, has higher fibrin
specificity than tPA (396). In the absence of fibrin, hardly any PLG
activation occurs (397,398). PLG activation by DSPA is barely
stimulated by D-dimer or the DDE complex, whereas these fibrin
fragments strongly enhance tPA activity (399). The crystal
structure of the catalytic domain of DSPA-

has been solved at

2.9 resolution (400) and shows strong similarity to tPA, which is


also active in the single-chain form. Experiments in animals
revealed that the thrombolytic potential of DSPA was at least equal
to, and the clearance rate approximately 10 times slower than,
that of tPA (401,402,403). The hope that high fibrin specificity
would result in lesser bleeding at comparable thrombolytic
potential was not fully realized (401), but this agent has several
promising features, long clearance time, and fibrin specificity,
which permits bolus administration and may be useful in stroke as
well as acute myocardial infarction (404).

INHIBITORS OF THE FIBRINOLYTIC


SYSTEM
Several inhibitors control and modulate the expression of
fibrinolytic activity. These act at the PLG activation step or on
plasmin (see Fig. 18-7), and they prevent excessive systemic
fibrinolytic activity. The principal inhibitor of tPA and uPA is PAI-1
(see Chapter 19), whereas 2 -antiplasmin inhibits plasmin. Both
are members of the serpin family, which also includes antithrombin
and heparin cofactor (see Chapter 13). The serpins all function in

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

48 de 138

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

the same way: They present to the target protease a reactive


center loop, the P 1 -P 1 of which mimics a substrate. A stable
inactive 1:1 complex is formed between protease and serpin,
lowering the effective protease concentration. Dissociation of the
covalent complex, which occurs at different rates for different
proteaseserpin pairs, results in a cleaved form of the serpin,
which lacks the short peptide from P 1 to the original C- terminus.
The serpins have marked specificity for particular proteases, but
the specificity is not absolute, so that other inhibitors present at
high concentration provide a back-up inhibitory system.

2 -Antiplasmin
2 -Antiplasmin, also called plasmin inhibitor, is the primary
fast-acting inhibitor of plasmin (405). It is a single-chain molecule
of M r of 70,000, comprising 452 residues, the additional mass
accounted for by approximately 13% carbohydrates on four
potential asparagine glycosylation sites. The plasma concentration
of 2 -antiplasmin is 70 mg per L (i.e., 1 M), approximately half
the concentration of PLG on a molar basis. It is synthesized in the
liver and consequently decreased in patients with advanced
impairment of hepatic function. The t 1/2 of the native inhibitor is
approximately 3 days, whereas the plasmin/ 2 -antiplasmin complex
is cleared with a t 1/ 2 of approximately 0.5 days (406). The cDNA
has been cloned (407), the organization of the gene determined
(408), and the amino acid sequence derived. The gene for
2 -antiplasmin (gene symbol PLI) has been localized to
chromosome 17p13 (409). The inhibitory activity of 2 -antiplasmin
resides in its reactive center loop, in which the P 1 -P 1 is ArgMet
(see Fig. 18-8); as with
P.350
other serpins, mutation of Arg to Ala destroys inhibitory function
(410). The rate of interaction of wild-type 2 -antiplasmin with
plasmin is very fast, with a rate constant of 4 10 7 per M per
second (411), comparable with the interaction of tPA and PAI-1.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

49 de 138

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

FIGURE 18-7. Activators and inhibitors of the blood


fibrinolytic system. scuPA, single-chain urinary plasminogen
activator; sctPA, single-chain tissue plasminogen activator;
HMWK, high-molecular-weight kininogen; z -AP
2 -antiplasmin; 2 M 2 -macroglobulin; C1-INH, complement C1
esterase inhibitor; HRG, histidine-rich glycoprotein; PAI-1,
plasminogen activator inhibitor 1.

Plasma 2 -antiplasmin can occur in four different forms, depending


on limited proteolysis at N- and C-termini. This processing appears
not to affect the inhibitory capacity of 2 -antiplasmin which resides
in the core of the protein. The N-terminus of 2 -antiplasmin was
originally identified as Asn (407), but it was later recognized that
there is also a proform with an additional 12 residues at the
N-terminus (412). Plasma 2 -antiplasmin was shown to have about
equal amounts of the two forms (413), whereas HepG2 cells
produced the Met-form, which, after addition to plasma, was
cleaved to the Asn form (414). An enzyme that removes the 12
residue N-terminal fragment has been purified and characterized

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

50 de 138

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

(415). The importance of the N-terminal processing lies in the


accessibility of Gln (at position 2 in the processed form); the
proform terminus blocks access to the cross-linking site (412).
Gln2 in 2 -antiplasmin is cross-linked by factor XIIIa to the A
chain of fibrinogen or fibrin, at Lys303 (416,417). Fibrin with
bound 2 -antiplasmin is markedly more resistant to lysis than
fibrin alone, and this observation was the basis of discovering the
first deficiency of 2 -antiplasmin (416). The cross-linking site can
be saturated, as elegantly shown by studies with wild-type
2 -antiplasmin and an active site Arg to Ala mutant (418).
Antibodies specific for cross-linked 2 AP greatly stimulated lysis
of fibrin (419).
Other studies focused on the C-terminus of 2 -antiplasmin which is
extended relative to other serpins by approximately 50 residues
(407). Two forms are detectable in normal human plasma (420).
The native, full-length form binds PLG, whereas a processed form
retains inhibitory capacity but cannot bind PLG (421). The two
forms were distinguished by two- dimensional
immunoelectrophoresis, with PLG incorporated in the gel in the
first dimension (422). The ratio of the PLG
P.351
binding to nonbinding form remains relatively constant in plasma
samples, at approximately 65:35, even in patients with advanced
liver cirrhosis (330).

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

51 de 138

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

FIGURE 18-8. Interaction of 2 -antiplasmin with plasmin and


with the chain of fibrinogen. For Lysplasmin, only the
high-affinity lysine binding site (LBS) 1 is indicated. Hdenotes the N-termini and -OH the C-termini of the three
proteins depicted. Lys residues in the C-terminus of
2 -antiplasmin interacts with the LBS 1 of plasmin, and the
N-terminal Gln2 is covalently cross-linked to Lys303 in the
chain of fibrin by factor XIIIa. tPA, tissue-type plasminogen
activator; UK, urokinase. (From Bachmann F. Fibrinolysis. In:
Verstraete M Vermylen J, Lijnen HR, et al., eds. Thrombosis
and haemostasis. Leuven: Leuven University Press, 1987:227,
with permission.)

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

52 de 138

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

The ability of 2 -antiplasmin to bind PLG interferes with the


fibrin-PLG interaction, because both involve LBS-1, as illustrated in
Figure 18-8. It is also crucial to the fate of plasmin formed on the
fibrin surface, to which both tPA and PLG bind by virtue of their
lysine-binding sites. Plasmin formed on fibrin is therefore relatively
protected from the action of 2 -antiplasmin (423). This concept
arose from experiments using lysine analogues, where 2 AP was
shown to be approximately 100 times less effective toward plasmin
in the presence of lysine analogues (424). The exact Lys residues
responsible for binding the C-terminal region of 2 -antiplasmin to
LBS-1 in PLG are not yet defined. Binding of recombinant
C-terminal peptide (55 residues) showed that binding to isolated
kringles was most affected by mutation of Lys452 but that other
internal Lys residues tether the kringles (425). A different study,
in which each Lys from position 429 to the C-terminal Lys452 was
separately mutated, indicated that Lys436 had the greatest effect;
their modeling suggested that Lys452 is too close to Phe448 for it
to be able to interact with the LBS (426).
The importance of 2 AP in stabilizing fibrin has been revealed in
several in vitro systems, including clot lysis (427,428). A mutant
of tPA that is relatively resistant to 2 AP was very efficient in
lysing both clots and model thrombi (429). Deficiency of 2 AP
results in delayed bleeding, but initial hemostasis is normal, and
this is characteristic of fibrinolytic defects (217). Inheritance was
autosomal recessive in a well-characterized Japanese family
(430,431). An interesting defect in 2 AP Enschede was explained
by the abnormal protein behaving as a substrate for plasmin rather
than as an inhibitor (432), as a result of the extended (by one Ala
residue) reactive center loop (433). Mice deficient in 2 AP were
essentially normal in many respects, but they did show enhanced
endogenous fibrinolytic activity without overt bleeding (434). In
experimental models, the lack of 2 AP had effects on liver
regeneration (435), platelet aggregation (436), and pulmonary
heart failure (437).

2 -Macroglobulin
2 -Macroglobulin is a major inhibitor of wide specificity and is the
backup inhibitor of several proteases, including plasmin (438). It

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

53 de 138

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

comes into play when the fibrinolytic system is fully activated and
when the capacity of 2 -antiplasmin is exhausted; it also forms
complexes with tPA and uPA (183). 2 -Macroglobulin is a
glycoprotein of M r 725,000 that consists of four identical chains of
M r 160,000 and contains approximately 8% carbohydrate (Table
18-1). Its plasma concentration is 2.5 g per L (3 M). The tetramer
is arranged as a pair of dimers, each consisting of two monomers
linked together via a disulfide bond. The complete sequence of this
1,451-residue molecule has been established and the gene locus
(gene symbol A2M) attributed to chromosome 12p13.3-p12.3
(439,440). The inhibitor contains two reactive sites. The sequence
Arg-Val-Gly-Phe-Tyr-Glu (681,682,683,684,685,686) acts as a bait
region that contains sites for proteases of different specificities
(441). Cleavage results in a conformational change and activates
the thioester site at positions 949952. Electron microscopic
studies showed that the catalytic domain of the rod-shaped
plasmin molecule is entrapped inside the 2 -macroglobulin cavity,
whereas its N-terminal kringle domains protrude outside one end
between the two armlike features of the transformed
2 -macroglobulin structure (442). It is noteworthy that proteases
inhibited by 2 -macroglobulin can still act on small peptide
substrates, but their position in the inhibitor cavity makes larger
targets nonaccessible.

PLASMINOGEN ACTIVATOR INHIBITOR 1


This is the principal inhibitor of both tPA and uPA and is discussed
in detail in Chapter 19; the emphasis in this chapter is therefore to
place it in the context of regulation of the PLG/plasmin system in
the vasculature. In addition to its inhibition of tPA and uPA, PAI-1
has a role in regulating cell adhesion processes relevant to tissue
remodeling, mediated by its binding to the adhesive glycoprotein
vitronectin, rather than by proteinase inhibition (443).
The 52-kDa glycoprotein consists of 379 residues. It is unusual
among the serpins in its tendency to lose activity through
spontaneous insertion of the reactive center loop into the body of
the molecule, forming an extra strand of -sheet A. This form was
termed latent because, after denaturation and refolding, activity
was regained (444); there are no physiologic conditions under

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

54 de 138

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

which this can occur. The noncovalent binding of vitronectin


stabilizes the active form of PAI-1 (445), which is present in
plasma at low concentrations, approximately 20 ng per mL
(446,447,448). Because it has a t 1/ 2 of less than 10 minutes, its
rate of synthesis must be high. The variation between normal
individuals is rather high, ranging from 1 to 40 ng per mL,
corresponding to 1 to 40 U per mL, the unit being the activity that
neutralizes one unit of tPA in 10 minutes.
PAI-1 is produced by most cultured cells, and early publications
identified it as a platelet protein (449), a product of
megakaryocytes (450), endothelial cells (451), or hepatocytes
(452); more recently, the production by adipocytes has been
stressed (453,454,455). The source of resting plasma PAI-1 is not
certain. Studies showing that normal endothelium or hepatocytes
do not express PAI-1 (456) point to adipocytes as the normal
source, whereas elevations in the acute-phase are likely to result
from hepatocytes (457). Platelets constitute the major pool of
circulating PAI-1. Although it is less active than plasma PAI-1,
platelet PAI-1 accounts for about half the circulating PAI-1 activity,
by virtue of the abundance of the protein in platelets (447). Large
amounts of platelet PAI-1 accumulate in thrombi (458,459), and
this directly correlates with their lysability (459).
PAI-1 inhibits tPA (tc and sc) and uPA (460). It does not inhibit
scuPA, which is largely inactive, but it does associate with scuPA
noncovalently, as discussed in the section on uPA (340). PAI-1 in
plasma is in excess over tPA; therefore, most of the tPA is present
in a complex with PAI-1. The second-order rate constants for the
interaction with tPA and uPA are greater than 10 7 M -1 S - 1 , a little
lower for sctPA (see Table 18-5).
Hereditary deficiency of PAI-1 in humans is quite rare but leads to
a lifelong bleeding disorder, presumably caused by premature lysis
of hemostatic plugs at sites of vascular trauma, and consistent
with this, bleeding occurs after a delay (461,462,463,464). The
oral administration of fibrinolytic inhibitors such as AMCA has been
effective in normalizing hemostatic function (463,464).
Mice with a disrupted PAI-1 gene demonstrate normal
development, fertility, and survival. The lack of inhibitor activity

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

55 de 138

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

induces a mild hyperfibrinolytic state and confers a greater


resistance to the prothrombotic effect of endotoxin injected into
the footpad (465). After arterial injury, wound healing in PAI - / mice was accelerated with rapid migration of smooth muscle cells
from the uninjured tissue (466). In an arterial thrombosis model,
time to development of an occlusive thrombus was approximately
twice as long in

-/-

than in

-/-

mice (467). In another model of

carotid artery injury, platelet-rich thrombi were generated and


residual thrombi evaluated 24 hours later. Partially or completely
occlusive thrombi were found in 34% of
mice (468).

-/-

mice and 65% of

- /-

TABLE 18-5 SECOND-ORDER RATE CONSTANTS FOR THE


INHIBITION OF TISSUE-TYPE PLASMINOGEN ACTIVATOR
AND URINARY-TYPE PLASMINOGEN ACTIVATOR BY
SERPINS

Second-order rate constant of


inhibition/M/s
sctPA

tctPA

uPA

PAI-1

10 7

3 10 7

5 10 7

PAI-2

10 3

2 10 5

10 6

Protease nexin

1.510 3

3 10 4

1.5 10 6

Protein C
inhibitor

10 3

10 3

C1-inhibitor

1.3

2 -Antiplasmin

13

[approximate,

[approximate,

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

56 de 138

equals]10 2

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

equals]10 2

sctPA, single-chain tissue-type plasminogen activator;


tctPA, two-chain tissue-type plasminogen activator; uPA,
urinary-type plasminogen activator; PAI-1, plasminogen
activator inhibitor-1; PAI-2, plasminogen activator
inhibitor-2; C1-inhibitor, complement 1 esterase inhibitor.
From Kruithof EKO. Inhibitors of plasminogen activators. In:
Kluft C, ed. Tissue-type plasminogen activator (tPA).
Physiological and clinical aspects, Vol I. Boca Raton, FL:
CRC Press, 1988:189210; Lobov S, Wilczynska M,
Bergstrom F, et al. Structural bases of the redox-dependent
conformational switch in the serpin PAI-2. J Mol Biol
2004;344:13591368; Huisman LGM, Van Griensven JMT,
Kluft C. On the role of C1-inhibitor as inhibitor of
tissue-type plasminogen activator in human plasma. Thromb
Haemost 1995;73:466471, with permission.

P.352
The opposite phenomenon, high plasma PAI-1, has been the
subject of many studies that examined potential links with
thrombosis. Mice overexpressing human PAI-1 in endothelium are
subject to age-dependent development of arterial thrombosis
(469). In humans, high circulating PAI-1 is clearly associated with
a variety of pathologies, including cardiovascular disease
(470,471,472) and cancer (473). Many studies have been
conducted to investigate causal significance, and it appears that
high PAI-1 does not independently predict disease, once other
factors like obesity, diabetes, and elevated triglycerides are taken
into account (474). A guanine insertion/deletion polymorphism at
position 675 in the PAI-1 promoter (475) is associated with
differences in circulating PAI-l (476). Despite promising early
findings (477), the current view is that the predictive power of
studying this polymorphism is low (474,478). In contrast to these
studies on subtle variation in plasma PAI-1, Gram-negative
septicemic patients have plasma PAI-1 concentraions that are

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

57 de 138

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

sometimes raised as much as 50-fold over normal, and this is


associated with high mortality (479).

Chemical Inhibitors of Plasminogen


Activator Inhibitor 1
Compounds that inhibit the synthesis or action of PAI-1 have been
sought because of the relation between elevated PAI-1 and
disease. They might be useful for enhancement of constitutive
fibrinolytic activity or as adjuvant drugs in the thrombolytic
treatment of thrombotic disorders. Gemfibrozil and clofibric acid
act at the level of PAI-1 synthesis, decreasing it by approximately
40% (480). Also acting at the level of PAI-1 synthesis, a butadiene
derivative T-686, administered to mice for 8 days before an
injection of lipopolysaccharide, protected mice from its lethal
effect in a dose-dependent fashion (481). In a rat thrombosis
model, it limited thrombus growth (482) and also attenuated
development of atherosclerotic lesions in rabbits subjected to a
high-cholesterol diet (483).
A 14amino acid peptide, corresponding to the PAI-1 reactive
center loop (residues 333 to 346), rapidly inhibited PAI-1 function
and prevented the formation of a complex between tPA and PAI-1.
It also enhanced considerably in vitro lysis of platelet-rich clots
(484). Sideroxylonal C, a phloroglucinol dimer isolated from the
flowers of Eucalyptus albens, inhibited PAI-1 in an in vitro assay
without having a considerable effect on tPA (485). AR-H029953XX
is a derivative of flufenamic acid, known to inhibit the C1-inhibitor.
It prevents complex formation of tPA with PAI-1 through binding to
Arg76 or Arg118, or both (486). A series of diketopiperazine-based
inhibitors of PAI-1, XR334, XR1853, and XR5082, inhibited the
action of PAI-1 on tPA and tcuPA (IC 50 : 5 to 80 M) and slowed the
time to blood vessel occlusion in a rat carotid artery injury model
(487). XR5118 binds to a region in PAI-1 (residues 110 to 145)
known to interact with tPA, and it stimulated endogenous lysis of
rabbit jugular vein thrombi and diminished thrombus growth (488).
XR5118 binds to PAI-1 at -sheet A, which causes an inactivating
conformational change (489).

Plasminogen Activator Inhibitor 2

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

58 de 138

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

The gene for PAI-2 (gene symbol PLANH2) spans 16.5 kb and is
located on chromosome 18q22.1 (490). The cDNA of 1.9 kb
encodes a protein of 415 residues. A segment of more than 5 kb
has been sequenced in the 5-flanking region and contains many
regulatory elements important for constitutive and for stimulated
gene expression (491,492).
PAI-2 was initially identified as a urokinase inhibitor in human
placenta (493); it later was purified from human placenta and from
the monocytoid cell line U-937 (494,495). PAI-2 is the
predominant PA inhibitor of squamous epithelia in epidermis,
esophagus, cornea, oral mucosa, tongue, and vagina (496). It is
incorporated into the cornified envelope during terminal
differentiation of the keratinocyte via transglutaminase-catalyzed
crosslinks (497). Monocytes are an important source of PAI-2
(498) and may increase fibrin stability on migration into thrombi,
especially because the PAI-2 is cross-linked to fibrin (499).
PAI-2 belongs to a subgroup of serpins called the
ovalbumin-related serpin family (ov-serpins). Members of this
subgroup lack a cleavable N-terminal signal sequence (490), and
accordingly much of the protein is intracellular. In addition to this
nonglycosylated form of apparent mass 47 kDa, some secreted,
glycosylated 60-kDa PAI-2 is found in cultured U937 cells (500).
All of the potential glycosylation sites on Asn75, Asn115, and
Asn339 appear to be used. Peripheral blood monocytes also secrete
some nonglycosylated PAI-2 by a mechanism distinct from the
secretion system for interleukin 1 and other nonclassically
secreted proteins (499,501). PAI-2 exhibits an unusually long
sequence between helices C and D, 33 residues that probably
protrude from the molecule. The crystal structure for PAI-2 is of a
form from which this loop was deleted (502). The loop contains
three glutamines in positions 83, 84, and 86 that are essential for
transglutaminase-mediated cross-linking of PAI-2 to trophoblast
membranes, the cornified envelope of epidermis (503), and fibrin
(499,504). Residues between 66 and 98 act as a protein-binding
domain, which binds annexins I, II, IV, and V, among others (505).
Purified PAI-2 polymerizes spontaneously at room temperature by
inserting the reactive center loop into the A-sheet of another
molecule, a loop-sheet polymerization that has some similarity to

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

59 de 138

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

that which occurs in the Z-mutant of 1 -antitrypsin (506). This


polymerization depends on disulfide bond formation between Cys
residues 37 (in the C-D loop) and 161, which can occur as a result
of the mobility of the C-D loop (507). Alternatively,
P.353
monomeric PAI-2, with a reduced Cys37, can form disulfide bonds
with other molecules such as vitronectin (508).
PAI-2 is an efficient inhibitor of tcuPA, with a second-order rate
constant of 10 6 M - 1 S -1 ; it reacts approximately five times more
slowly with tctPA and quite poorly with sctPA (Table 18-5)
(460,495). In normal plasma, there are no measurable levels of
PAI-2, but high levels are found in pregnancy, with a steady rise
until approximately 33 weeks, to reach approximately 250 ng per
mL (509). Lesser increases during pregnancy were observed during
intrauterine growth retardation (510,511,512). PAI-2 might
therefore be a marker for decreased placental function (511,512).
Different forms of PAI-2 occur in plasma in pregnancy (513), and it
is not clear whether these represent PAI-2 disulfide-bonded or
cross-linked to other proteins.
PAI-2 is also found in the plasma of patients with acute
myeloblastic leukemia of the M 4 and the M 5 type (514), and in
patients with sepsis, in whom PAI-1 is also increased (515).
Speculation remains on the real function of PAI-2. Its intracellular
location suggests roles other than inhibition of PA, and it was
found to protect cells from the rapid cytopathic effects of
alphavirus infection (516). It also acts intracellularly as a
retinoblastoma protein (Rb)-binding protein (517). Its local activity
appears to be relevant to a number of cancers (518,519). Mice
with a disrupted PAI-2 gene did not present any phenotypic
abnormalities (520); it will be interesting to see if appropriate
inflammatory challenges reveal a phenotype.

C1-Inhibitor
C1-inhibitor is a highly glycosylated 105-kDa serpin, consisting of
478 residues (Tables 18-1 and 18-2). The 16 kb gene is located on
chromosome 11q12-q13.1 (521,522,523). It inhibits the activated
subcomponents, C1r and C1s, of complement C1, and factors XIIa

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

60 de 138

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

and XIa, plasma kallikrein, and plasmin. Its high plasma


concentration (1.7 M) gives it the capacity to inhibit tPA when it
is present in excess over PAI-1 (138,330,524). By virtue of its
inhibitory effect on components of the contact system, it probably
plays a role in controlling contact phasedependent fibrinolysis and
the conversion of scuPA to uPA.

Histidine-Rich Glycoprotein
Histidine-rich glycoprotein (HRG) is named for the region between
residues 330 and 389, which are similar to the histidine-rich
regions in human and bovine kininogen (525). The 75-kDa protein
has 507 residues and approximately 14% of carbohydrate. HRG
competitively inhibits PLG associations by binding LBS-1 of PLG (K d
1 M) It is suggested, on the basis of their relative concentrations,
that PLG (2 M) and HRG (1.5 M) would circulate predominantly
in a 1:1 reversible complex (526), which would decrease the PLG
available for binding to fibrin. However, 2 -antiplasmin had a
greater effect than HRG on binding of PLG to fibrin (527). It has
recently been proposed that HRG tethers PLG to the cell surface,
which would enhance the cells' migratory potential (528). Only one
family with congenital heterozygous HRG deficiency has been
described. The proband, but none of the other five family members
with low HRG levels, had thrombotic disease (529).

Thrombin-Activatable Fibrinolysis
Inhibitor
The generation of C-terminal lysyl residues is an important
feedback mechanism for the binding of PLG to fibrin and the
enhancement of fibrinolysis. These residues can be removed by
TAFIa (EC 3.4.17.20; synonyms: carboxypeptidase B, U, or R),
further regulating fibrinolysis (530), as discussed fully in Chapter
20. The zymogen TAFI is activated by thrombin/ thrombomodulin
and functionally connects the fibrinolytic and coagulant pathways
(531).
TAFI is primarily produced by the liver and circulates at
approximately 75 nM, but the normal variation is wide (530,532).
The more important variable is the circulating active enzyme,
which has a very short t 1/ 2 and only a small proportion has to be

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

61 de 138

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

activated for full function (533). The function of TAFIa was shown
originally in clot lysis assays; potato tuber carboxypeptidase
inhibitor relieves the inhibition (531,534). The effect of TAFIa has
also been shown in model thrombi made from whole blood (535)
and in animal studies, where TAFIa activity increased the time
taken to restore blood flow in a canine coronary artery thrombosis
model (536). Inhibitors of TAFIa enhanced endogenous fibrinolysis
and also thrombolysis by tPA (537,538). The homozygous-deficient
mouse shows enhanced plasma clot lysis but is comparable to
wild-type mouse in many thrombus models, mostly those in which
time taken to occlusion rather than thrombus resolution is
assessed (539).
There are some reports of elevated TAFI as a mild risk factor for
thrombosis, and it is elevated in inflammation, correlating with
other acute-phase markers (540). Several polymorphisms in the
TAFI gene have been reported. These seem to explain the wide
normal range of concentrations but do not correlate strongly with
disease (541). Some methods show bias toward particular
polymorphic variants of TAFI (542,543); associations with disease
therefore need to be interpreted with caution.
TAFIa activity appears to be the reason for increased fibrinolysis in
hemophilia, where thrombin generation may be sufficient for fibrin
formation but suboptimal for TAFI activation. Most thrombin is
formed after clot formation, mainly via the intrinsic pathway by
activation of factor XI by thrombin. This leads to a positive
feedback mechanism and optimal activation of pro-TAFI (544).
Factor IXdeficient plasma clots lysed prematurely, but the defect
could be reversed by the addition of factor IX or of
thrombomodulin to the plasma (545). Similarly, addition of TAFI,
thrombomodulin, or factor VIII to hemophilia A plasma restored
normal fibrinolysis (546). Consistent with this, incorporation of
antifactor XI antibodies or inhibition of TAFI in a rabbit model
resulted in an almost twofold increase of endogenous thrombolytic
activity (547).

Lipoprotein (a)
Apolipoprotein (a), a component of lipoprotein (a) [Lp(a)], is
linked by a disulfide bond to apolipoprotein B-100. Apo(a) has a

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

62 de 138

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

wide size distribution, which depends on the number of kringle


repeats; these are similar to kringle 4 of PLG (548). The total mass
of Lp(a) therefore varies between 800 and 1,300 kDa. There is also
a wide range of plasma Lp(a) concentrations, from 10 to 1,000 mg
per L, and a skewed distribution, median value 50 mg per L.
Besides apo(a), a gene cluster on chromosome 6q26-27 also
contains PLG and two apo(a)-related genes (11,549,550).
Lp(a) binds to fibrin, extracellular matrix, platelets, and cells by
its kringles; predictably, binding is inhibited by lysine and ACA.
Lp(a) therefore competes with PLG and tPA for binding to fibrin
and may thus exert an antifibrinolytic effect (551,552,553). PLG
binding and activation on platelets (554) as well as on fibrin
(555,556,557,558) is attenuated by Lp(a). Transgenic mice
expressing the human apo(a) gene resolved experimental thrombi
poorly when treated with tPA (559). These observations are
consistent with the association of high Lp(a) with coronary artery
disease (560,561,562). It is now clear that Lp(a) size is a key
factor; only the smaller forms bind with high
P.354
affinity to fibrin (563,564,565). Individuals with the combination of
high Lp(a) levels and predominance of the small isoform showed
the lowest fibrinolytic activity (565).

BALANCE OF THE SYSTEM


The fibrinolytic system is normally kept quiescent as a result of
two factors: the relative abundance of inhibitors over proteases
and the requirement for fibrin or cell surfaces for initiation of
activity. In normal individuals, steady-state levels of PLG in
plasma are rather stable, whereas both the PA and their inhibitors
are subject to change, suggesting that regulation of fibrinolytic
activity occurs mainly via up- and downregulation of the
expression of the PA and their inhibitors. This general statement
relates both to the relative concentrations of PA, many orders of
magnitude lower than the PLG concentration, and to the very short
half-lives of PA compared to PLG, minutes rather than days (Table
18-2).
The fibrinolytic system is geared to remove fibrin from the

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

63 de 138

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

circulation in a controlled way, and therefore to prevent excessive


fibrin accumulation. Circulating PLG is not normally converted to
plasmin, which would cause generalized proteolysis and
degradation of clotting factors, particularly factor V, factor VIII,
VWF, and platelet glycoproteins. Instead, fibrin acts as a focus for
the generation of activity, which is relatively protected from
inhibitors. Although fibrin deposition plays an important role in
hemostasis, wound healing, and inflammatory processes, fibrin
formation must always be limited in time, and fibrin should be
removed once it has fulfilled its role. The primary role of the
PLGplasmin system is not to lyse large formed blood clots, but to
prevent excessive build-up of fibrin formation. That said, the
system has been harnessed usefully for therapy, mainly by
administering activators. Recently there has been a resurgence of
interest in therapeutic use of plasmin (566).
Fibrin acts to assemble the fibrinolytic components, which is
central to the control of the system. At the earliest stages of fibrin
formation, tPA and PLG bind to the forming fibrin strands. Once
small amounts of tPA and PLG are bound to fibrin in the form of a
ternary complex, the catalytic efficiency of tPA for PLG is several
hundred times higher than it is in the absence of fibrin. Plasmin
generation causes proteolytic cleavage of fibrin that starts at the
C-terminal portion of the chain of fibrin and produces new
C-terminal lysyl residues. Partially digested fibrin binds up to 10
times more GluPLG than undegraded fibrin. scuPA binds to and
appears to activate selectively GluPLG bound to C-terminal
lysines in the partially degraded fibrin. Trace amounts of plasmin
also activate scuPA to uPA, and there is reciprocal activation of
PLG to plasmin.
In healthy individuals, there is a molar excess of active PAI-1 in
plasma over tPA (Table 18-2). Consequently, most of the tPA in
plasma is in the form of an inactive complex with PAI-1 (241).
Small amounts of free tPA occur in the circulation as a result of the
high synthetic rate of both proteins and the finite time it takes for
complex formation to occur. Even though PAI-1 in plasma is kept
in its active conformation by its interaction with vitronectin, its
tendency to lose activity does increase the possibility of plasma
tPA expressing some of its activity. Normally, tPA activity is

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

64 de 138

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

detectable at approximately 0.1 to 0.5 U per mL of plasma, even


when blood is collected in an acidified anticoagulant to prevent
complexing of any free tPA by PAI-1 (567). The maximum activity
normally seen is equivalent to 1 g per L; tPA has a specific
activity of 500,000 to 700,000 U per mg (568). This low tPA
activity is generally close to the borderline for detection, which is
why procedures like euglobulin preparation, which removes most
plasma inhibitors, but notably only about half the PAI-1 (447), are
used. Even after the levels of free tPA increase severalfold, after
strenuous exercise, venous occlusion, or intravenous injection of
DDAVP to healthy volunteers, no circulating plasmin forms because
PLG activation does not occur at these still physiologic tPA
concentrations. This does not apply during the treatment of acute
myocardial infarction with recombinant tPA, during which
circulating plasma concentrations of tPA are achieved that are up
to 1,000-fold higher than those observed in normal plasma.
Changes in the balance of tPA and PAI-1 are a clear example of
how fibrinolytic activity can be initiated. Indeed, these components
of the system fluctuate remarkably during a 24-hour period (569).
tPA activity is lowest at night and in the early morning and then
increases up to threefold during the day. PAI-1 antigen and
activity are highest in the early morning hours and then gradually
decline during the day, reaching their lowest level in the early
afternoon. Consequently, if a drug (or placebo) is given at 8 AM,
the activity of the fibrinolytic system invariably is higher a few
hours later, and only trials in which a comparison to placebo is
made are valid. Various factors may interfere with the normal
circadian rhythm (327).
There is a large literature on the predictive value of tPA antigen
and PAI-1 as risk factors for acute myocardial infarction, stroke,
and peripheral vascular disease (472,477,478,570). High tPA
antigen is found to be associated with disease in a number of
studies, and this probably reflects the tPA/PAI-1 complex
(241,571).
It is generally assumed that PLG, present in plasma at
approximately 2 M, is not rate-limiting for fibrinolysis. Even
though the concentration is so much higher than the activators

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

65 de 138

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

(Table 18-2), it is indeed possible for it to become a limiting


factor. The need to consider local PLG concentration, especially
when the activators are used therapeutically, has been clearly
shown (572). The effect on variation in available PLG is highlighted
in studies on TAFIa (573). The effect of TAFIa on tPA-initiated lysis
is usually stressed, in line with tPA's known fibrin dependence, but
TAFIa also has strong effects on uPA and scuPA-induced lysis,
pointing to its primary effect being on PLG binding (535).
Global circulating (a rare observation, unless thrombolytic agents
are administered to patients at high doses) and local fibrinolytic
activity are modulated by serpins. tPA and uPA are efficiently
inhibited by PAI-1. More than 90% of blood PAI-1 is located in the
platelets and promptly released on platelet activation. Although
most of the platelet PAI-1 is inactive, there is still enough active
PAI-1 in platelets to stabilize platelet-rich thrombi, as occur in the
arterial circulation in which the blood pressure is high and
premature lysis of a hemostatic plug is undesirable. Plasmin bound
to fibrin is partly protected from inactivation by 2 -antiplasmin
This assures that fibrinolysis proceeds on the surface of a clot.
However, free plasmin that spills over into the general circulation
is rapidly inactivated by 2 -antiplasmin After massive activation of
the fibrinolytic system has taken place, such as in thrombolytic
therapy, the amount of 2 -antiplasmin does not suffice to inhibit
all circulating plasmin, and 2 -macroglobulin acts as a scavenger
protease inhibitor.
The fibrinolytic system is a highly modulated enzyme system, and
nature has taken many steps to limit its action in time and space
but has also provided the necessary feedback loops that enhance
the fibrinolytic system on the local level.

References
1. Davidson CJ, Tuddenham EG, McVey JH. 450 million years of
hemostasis. J Thromb Haemost 2003;1:14871494.
2. Patthy L. Evolutionary assembly of blood coagulation
proteins. Semin Thromb Hemost 1990;16:245259.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

66 de 138

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

P.355
3. Madison EL. Probing structure-function relationships of
tissue-type plasminogen activator by site-specific mutagenesis.
Fibrinolysis 1994;8(Suppl. 1):221236.
4. Collen D, De Maeyer L. Molecular biology of human
plasminogen. I. Physicochemical properties and
microheterogeneity. Thromb Diath Haemorrh 1975;34:396402.
5. Raum D, Marcus D, Alper CA, et al. Synthesis of human
plasminogen by the liver. Science 1980;208:10361037.
6. Highsmith RF, Kline DL. Kidney: primary source of
plasminogen after acute depletion in the cat. Science
1971;174:141142.
7. Twining SS, Wilson PM, Ngamkitidechakul C. Extrahepatic
synthesis of plasminogen in the human cornea is up-regulated
by interleukins-1 and -1 . Biochem J 1999;339:705712.
8. Zhang L, Seiffert D, Fowler BJ, et al. Plasminogen has a
broad extrahepatic distribution. Thromb Haemost
2002;87:493501.
9. Murray JC, Buetow KH, Donovan M, et al. Linkage
disequilibrium of plasminogen polymorphism and assignment of
the gene to human chromosome 6q26-6q27. Am J Hum Genet
1987;40:338350.
10. Frank SL, Klisak I, Sparkes RS, et al. A gene homologous to
plasminogen located on human chromosome 2q11-p11.
Genomics 1989;4:449451.
11. Ichinose A. Multiple members of the
plasminogen-apolipoprotein(a) gene family associated with

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

67 de 138

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

thrombosis. Biochemistry 1992;31:31133118.


12. Sottrup-Jensen L, Claeys H, Zajdel M, et al. The primary
structure of human plasminogen: isolation of two lysine-binding
fragments and one mini-plasminogen (MW 38,000) by
elastase-catalyzed-specific limited proteolysis. In: Davidson JF,
Rowan RM, Samama MM et al., eds. Progress in chemical
fibrinolysis and thrombolysis, Vol. 3. New York: Raven Press,
1978: 191209.
13. Petersen TE, Martzen MR, Ichinose A, et al. Characterization
of the gene for human plasminogen, a key proenzyme in the
fibrinolytic system. J Biol Chem 1990;265:61046111.
14. Degen SJF, Bell SM, Schaefer LA, et al. Characterization of
the cDNA coding for mouse plasminogen and localization of the
gene to mouse chromosome 17. Genomics 1990;8:4961.
15. Wiman B, Walln P. The specific interaction between
plasminogen and fibrin. A physiological role of the lysine
binding site in plasminogen. Thromb Res 1977;10:213222.
16. Thorsen S, Clemmensen I, Sottrup-Jensen L, et al.
Adsorption to fibrin of native fragments of known primary
structure from human plasminogen. Biochim Biophys Acta
1981;668:377387.
17. Wiman B, Lijnen HR, Collen D. On the specific interaction
between the lysine-binding sites in plasmin and complementary
sites in 2 -antiplasmin and in fibrinogen. Biochim Biophys Acta
1979;579:142154.
18. Clemmensen I, Petersen LC, Kluft C. Purification and
characterization of a novel, oligomeric, plasminogen kringle 4
binding protein from human plasma: tetranectin. Eur J Biochem
1986;156:327333.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

68 de 138

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

19. Lijnen HR, Hoylaerts M, Collen D. Isolation and


characterization of a human plasma protein with affinity for the
lysine binding sites in plasminogen. J Biol Chem
1980;255:1021410222.
20. Kelm RJ Jr, Swords NA, Orfeo T, et al. Osteonectin in matrix
remodeling. A plasminogen-osteonectin-collagen complex. J Biol
Chem 1994;269: 3014730153.
21. Selim TE, Ghoneim HR, Uknis AB, et al.
High-molecular-mass and low-molecular-mass kininogens block
plasmin-induced platelet aggregation by forming a complex with
kringle 5 of plasminogen/plasmin. Eur J Biochem
1997;250:532538.
22. Silverstein RL, Leung LLK, Harpel PC, et al. Complex
formation of platelet thrombospondin with plasminogen:
modulation of activation by tissue activator. J Clin Invest
1984;74:16251633.
23. Deutsch DG, Mertz ET. Plasminogen: purification from
human plasma by affinity chromatography. Science
1970;170:10951096.
24. Rejante MR, Llins M. Solution structure of the
-animohexanoic acid complex of human plasminogen kringle 1.
Eur J Biochem 1994;221:939949.
25. Marti DN, Schaller J, Llins M. Solution structure and
dynamics of the plasminogen kringle 2-AMCHA complex:
3 1 -helix in homologous domains. Biochemistry
1999;38:1574115755.
26. Shndel S, Hu CK, Marti D, et al. Recombinant gene
expression and

H-NMR characteristics of the kringle (2 + 3)

supermodule: spectroscopic functional individuality of

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

69 de 138

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

plasminogen kringle domains. Biochemistry


1996;35:23572364.
27. Atkinson RA, Williams RJP. Solution structure of the kringle
4 domain from human plasminogen by nuclear magnetic
resonance spectroscopy and distance geometry. J Mol Biol
1990;212:541552.
28. Wu T-P, Padmanabhan KP, Tulinsky A. The structure of
recombinant plasminogen kringle 1 and the fibrin binding site.
Blood Coagul Fibrinolysis 1994;5:157166.
29. Mathews II, Vanderhoff-Hanaver P, Castellino FJ, et al.
Crystal structures of the recombinant kringle 1 domain of
human plasminogen in complexes with the ligands
-animocaproic acid and
trans-4-(aminomethyl)cyclohexane-1-carboxylic acid.
Biochemistry 1996;35:25672576.
30. Wu T-P, Padmanabhan K, Tulinsky A, et al. The refined
structure of the -animocaproic acid complex of human
plasminogen kringle 4. Biochemistry 1991;30:1058910594.
31. Chang Y, Mochalkin I, McCance SG, et al. Structure and
ligand binding determinants of the recombinant kringle 5
domain of human plasminogen. Biochemistry
1998;37:32583271.
32. Lerch PG, Rickli EE, Lergier W, et al. Localization of
individual lysine-binding regions in human plasminogen and
investigations on their complex-forming properties. Eur J
Biochem 1980;107:713.
33. Menhart N, Sehl LC, Kelley RF, et al. Construction,
expression, and purification of recombinant kringle 1 of human
plasminogen and analysis of its interaction with omega-amino

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

70 de 138

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

acids. Biochemistry 1991;30:19481957.


34. Novokhatny VV, Matsuka YV, Kudinov SA. Analysis of ligand
binding to kringles 4 and 5 fragments from human plasminogen.
Thromb Res 1989; 53:243252.
35. McCance SG, Menhart N, Castellino FJ. Amino acid residues
of the kringle-4 and kringle-5 domains of human plasminogen
that stabilize their interactions with -amino acid ligands. J Biol
Chem 1994;269:3240532410.
36. Marti D, Schaller J, Ochensberger B, et al. Expression,
purification and characterization of the recombinant kringle 2
and kringle 3 domains of human plasminogen and analysis of
their binding affinity for -animocarboxylic acids. Eur J
Biochem 1994;219:455462.
37. Christensen U, Sottrup-Jensen L, Magnusson S, et al.
Enzymic properties of the neo-plasmin-val-442 (miniplasmin).
Biochim Biophys Acta 1979; 567:472481.
38. Komorowicz E, Kolev K, Machovich R. Fibrinolysis with
des-kringle derivatives of plasmin and its modulation by plasma
protease inhibitors. Biochemistry 1998;37:91129118.
39. Lasters I, Van Herzeele N, Lijnen HR, et al. Enzymatic
properties of phage-displayed fragments of human plasminogen.
Eur J Biochem 1997; 244:946952.
40. Thewes T, Constantine K, Byeon IL, et al. Ligand
interactions with the kringle 5 domain of plasminogen. A study
by NMR spectroscopy. J Biol Chem 1990;265:39063915.
41. Motta A, Laursen RA, Llins M, et al. Complete assignment
of the aromatic magnetic resonance spectrum of the kringle 1
domain from human plasminogen: structure of the

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

71 de 138

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

ligand-binding site. Biochemistry 1987;26: 38273836.


42. Tulinsky A. The structures of domains of blood proteins.
Thromb Haemost 1991;66:1631.
43. Peisach E, Wang J, de los Santos T, et al. Crystal structure
of the proenzyme domain of plasminogen. Biochemistry
1999;38:1118011188.
44. Wang X, Terzyan S, Tang J, et al. Human plasminogen
catalytic domain undergoes an unusual conformational change
upon activation. J Mol Biol 2000;295:903914.
45. Walln P, Wiman B. Characterization of human plasminogen.
I. On the relationship between different molecular forms of
plasminogen demonstrated in plasma and found in purified
preparations. Biochim Biophys Acta 1970;221:2030.
46. Walln P, Wiman B. Characterization of human plasminogen.
II. Separation and partial characterization of different molecular
forms of human plasminogen. Biochim Biophys Acta
1972;257:122134.
47. Violand BN, Castellino FJ. Mechanism of the
urokinase-catalysed activation of human plasminogen. J Biol
Chem 1976;251:39063912.
48. Holvoet P, Lijnen HR, Collen D. A monoclonal antibody
specific for Lys-plasminogen. Application to the study of the
activation pathways of plasminogen in vivo. J Biol Chem
1985;260:1210612111.
49. Violand BN, Byrne R, Castellino FJ. The effect of -,-amino
acids on human plasminogen structure and activation. J Biol
Chem 1978;253: 53955401.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

72 de 138

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

50. Urano T, Chibber BAK, Castellino FJ. The reciprocal effects


of -aminohexanoic acid and chloride ion on the activation of
human [Glu 1 ] plasminogen by human urokinase. Proc Natl Acad
Sci U S A 1987;84: 40314036.
51. An SSA, Carreo C, Marti DN, et al. Lysine-50 is a likely site
for anchoring the plasminogen N-terminal peptide to
lysine-binding kringles. Protein Sci 1998;7:19601969.
52. Cockell CS, Marshall JM, Dawson KM, et al. Evidence that
the conformation of unliganded human plasminogen is
maintained via an intramolecular interaction between the
lysine-binding site of kringle 5 and the N-terminal peptide.
Biochem J 1998;333:99105.
53. Sjholm I, Wiman B, Walln P. Studies on the
conformational changes of plasminogen induced during
activation to plasmin by 6-aminohexanoic acid. Eur J Biochem
1973;39:471479.
54. Castellino FJ. Biochemistry of human plasminogen. Semin
Thromb Hemost 1984;10:1823.
55. Mangel WF, Lin B, Ramakrishnan V. Characterization of an
extremely large, ligand-induced conformational change in
plasminogen. Science 1990; 248:6973.
56. Stack S, Gonzales-Gronow M, Pizzo SV. The effect of
divalent cations of the conformation and function of human
plasminogen. Arch Biochem Biophys 1991;284:5862.
57. Thorsen S. Differences in the binding to fibrin of native
plasminogen and plasminogen modified by proteolytic
degradation. Influence of -aminocarboxylic carboxylic acids.
Biochim Biophys Acta 1972;393:5565.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

73 de 138

58. Rejante M, Elliott BW Jr, Llins M. A

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

H-NMR study of

plasminogen kringle 4 interactions with intact and partially


digested fibrinogen. Fibrinolysis 1991;5:8792.
59. Wu HL, Chang BI, Wu DH, et al. Interaction of plasminogen
and fibrin in plasminogen activation. J Biol Chem
1990;265:1965819664.
60. Lucas MA, Fretto LJ, McKee PA. The binding of human
plasminogen to fibrin and fibrinogen. J Biol Chem
1983;258:42494256.
P.356
61. Suenson E, Ltzen O, Thorsen S. Initial
plasmin-degradation of fibrin as the basis of a positive feedback
mechanism in fibrinolysis. Eur J Biochem 1984;140:513522.
62. Harpel PC, Chang T, Verderber E. Tissue plasminogen
activator and urokinase mediate the binding of Glu-plasminogen
to plasma fibrin I. Evidence for new binding sites in
plasmin-degraded fibrin I. J Biol Chem 1985;260:44324440.
63. Tran-Thang Ch, Kruithof EKO, Atkinson J, et al.
High-affinity binding sites for human Glu-plasminogen unveiled
by limited plasmic degradation of human fibrin. Eur J Biochem
1986;160:599604.
64. Fleury V, Angls-Cano E. Characterization of the binding of
plasminogen to fibrin surfaces: the role of carboxy-terminal
lysines. Biochemistry 1991;30:76307638.
65. Garman AM, Smith RAG. The binding of plasminogen to
fibrin: evidence for plasminogen-bridging. Thromb Res
1982;27:311320.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

74 de 138

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

66. Petersen LC, Suenson E. Effect of plasminogen and


tissue-type plasminogen activator on fibrin gel structure.
Fibrinolysis 1990;5:5159.
67. Crowe JD, Sievwright IK, Auld GC, et al. Candida albicans
binds human plasminogen: identification of eight
plasminogen-binding proteins. Mol Microbiol
2003;47:16371651.
68. Miles LA, Dahlberg CM, Plescia J, et al. Role of cell-surface
lysines in plasminogen binding to cells: identification of
-enolase as a candidate plasminogen receptor. Biochemistry
1991;30:16821691.
69. Cesarman GM, Guevara CA, Hajjar KA. An endothelial cell
receptor for plasminogen/tissue plasminogen activator (t-PA).
Annexin II-mediated enhancement of t-PA-dependent
plasminogen activation. J Biol Chem 1994; 269:2119821203.
70. Fischer MB, Roeckl C, Parizek P, et al. Binding of
disease-associated prion protein to plasminogen. Nature
2000;408:479483.
71. Ellis V. Plasminogen activation at the cell surface. Curr Top
Dev Biol 2003; 54:263312.
72. O'Reilly MS, Holmgren L, Shing Y, et al. Angiostatin: a
novel angiogenesis inhibitor that mediates the suppression of
metastases by a Lewis lung carcinoma. Cell 1994;79:315328.
73. Stathakis P, Fitzgerald M, Matthias LJ, et al. Generation of
angiostatin by reduction and proteolysis of plasmin. Catalysis
by a plasmin reductase secreted by cultured cells. J Biol Chem
1997;272:2064120645.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

75 de 138

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

74. Gately S, Twardowski P, Stack MS, et al. The mechanism of


cancer-mediated conversion of plasminogen to the angiogenesis
inhibitor angiostatin. Proc Natl Acad Sci U S A
1997;94:1086810872.
75. Lay AJ, Jiang XM, Kisker O, et al. Phosphoglycerate kinase
acts in tumour angiogenesis as a disulphide reductase. Nature
2000;408:869873.
76. Kwon M, Caplan JF, Filipenko NR, et al. Identification of
annexin II heterotetramer as a plasmin reductase. J Biol Chem
2002;277:1090310911.
77. Cornelius LA, Nehring LC, Harding E, et al. Matrix
metalloproteinases generate angiostatin: effects on
neovascularization. J Immunol 1998;161: 68456852.
78. Lijnen HR, Van Hoef B, Ugwu F, et al. Specific proteolysis of
human plasminogen by a 24 kDa endopeptidase from a novel
Chryseobacterium sp. Biochemistry 2000;39:479488.
79. Cao YH, Ji RW, Davidson D, et al. Kringle domains of human
angiostatin. Characterization of the anti-proliferative activity on
endothelial cells. J Biol Chem 1996;271:2946129467.
80. Joe Y-A, Hong Y-K, Chung D-S, et al. Inhibition of human
malignant glioma growth in vivo by human recombinant
plasminogen kringles 1-3. Int J Cancer 1999;82:694699.
81. MacDonald NJ, Murad AC, Fogler WE, et al. The
tumor-suppressing activity of angiostatin protein resides within
kringles 1 to 3. Biochem Biophys Res Commun
1999;264:469477.
82. Barendsz-Janson AF, Griffioen AW, Muller AD, et al. In vitro

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

76 de 138

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

tumor angiogenesis assays: plasminogen lysine binding site 1


inhibits in vitro tumor-induced angiogenesis. J Vasc Res
1998;35:109114.
83. Ji W-R, Barrientos LG, Llins M, et al. Selective inhibition
by kringle 5 of human plasminogen on endothelial cell
migration, an important process in angiogenesis. Biochem
Biophys Res Commun 1998;247:414419.
84. Ji W-R, Castellino FJ, Chang Y, et al. Characterization of
kringle domains of angiostatin as antagonists of endothelial cell
migration, an important process in angiogenesis. FASEB J
1998;12:17311738.
85. Lucas R, Holmgren L, Garcia I, et al. Multiple forms of
angiostatin induce apoptosis in endothelial cells. Blood
1998;92:47304741.
86. Walter JJ, Sane DC. Angiostatin binds to smooth muscle
cells in the coronary artery and inhibits smooth muscle cell
proliferation and migration in vitro. Arterioscler Thromb Vasc
Biol 1999;19:20412048.
87. Colman RW. The contact system and angiogenesis: potential
for therapeutic control of malignancy. Semin Thromb Hemost
2004;30:4561.
88. Kim JS, Chang JH, Yu HK, et al. Inhibition of angiogenesis
and angiogenesis-dependent tumor growth by the cryptic
kringle fragments of human apolipoprotein(a). J Biol Chem
2003;278:2900029008.
89. Hayes ML, Castellino FJ. Carbohydrate of the human
plasminogen variants. III. Structure of the O-glycosidically
linked oligosaccharide unit. J Biol Chem 1979;254:87778780.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

77 de 138

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

90. Hayes ML, Castellino FJ. Carbohydrate of the human


plasminogen variants. I. Carbohydrate composition,
glycopeptide isolation, and characterization. J Biol Chem
1979;254:87688771.
91. Summaria L, Arzadon L, Bernabe P, et al. Studies on the
isolation of the multiple molecular forms of human plasminogen
and plasmin by isoelectric focusing methods. J Biol Chem
1972;247:46914702.
92. Pirie-Shepherd SR, Jett EA, Andon NL, et al. Sialic acid
content of plasminogen 2 glycoforms as a regulator of
fibrinolytic activity. Isolation, carbohydrate analysis, and
kinetic characterization of six glycoforms of plasminogen 2. J
Biol Chem 1995;270:58775881.
93. Mlgaard L, Ponting CP, Christensen U. Glycosylation at
Asn-289 facilitates the ligand-induced conformational changes
of human Glu-plasminogen. FEBS Lett 1997;405:363368.
94. Gonzalez-Gronow M, Grenett HE, Fuller GM, et al. The role
of carbohydrate in the function of human plasminogen:
comparison of the protein obtained from molecular cloning and
expression in Escherichia coli and COS cells. Biochim Biophys
Acta 1990;1039:269276.
95. Hatton MWC, Southward S, Ross-Ouellet B. Catabolism of
plasminogen glycoforms I and II in rabbits: relationship to
plasminogen synthesis by the rabbit liver in vitro. Metabolism
1994;43:14301437.
96. Hatton MW, Day S, Ross B, et al. Plasminogen II
accumulates five times faster than plasminogen I at the site of
a balloon de-endothelializing injury in vivo to the rabbit aorta:
comparison with other hemostatic proteins. J Lab Clin Med
1999;134:260266.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

78 de 138

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

97. Pirie-Shepherd SR, Stevens RD, Andon NL, et al. Evidence


for a novel O-linked sialylated trisaccharide on Ser-248 of
human plasminogen 2. J Biol Chem 1997;272:74087411.
98. Hortin GL. Isolation of glycopeptides containing O-linked
oligosaccharides by lectin affinity chromatography on
jacalin-agarose. Anal Biochem 1990;191:262267.
99. Wang H, Prorok M, Bretthauer RK, et al. Serine-578 is a
major phosphorylation locus in human plasma plasminogen.
Biochemistry 1997;36: 81008106.
100. Skoda U, Bertrams J, Dykes D, et al. Proposal for the
nomenclature of human plasminogen (PLG) polymorphism. Vox
Sang 1986;51:244248.
101. Tefs K, Georgieva M, Seregard S, et al. Characterization of
plasminogen variants in healthy subjects and plasminogen
mutants in patients with inherited plasminogen deficiency by
isoelectric focusing gel electrophoresis. Thromb Haemost
2004;92:352357.
102. Yamaguchi M, Doi S, Yoshimura M. Plasminogen
phenotypes in a Japanese population. Four new variants
including one with a functional defect. Hum Hered
1989;39:356360.
103. Aoki N, Moroi M, Sakata Y. Abnormal plasminogen: a
hereditary molecular abnormality found in a patient with
recurrent thrombosis. J Clin Invest 1978;61:1186.
104. Schuster V, Zeitler P, Seregard S, et al. Homozygous and
compound-heterozygous type I plasminogen deficiency is a
common cause of ligneous conjunctivitis. Thromb Haemost
2001;85:10041010.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

79 de 138

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

105. Schuster V, Mingers AM, Seidenspinner S, et al.


Homozygous mutations in the plasminogen gene of two
unrelated girls with ligneous conjunctivitis. Blood
1997;90:958966.
106. Schuster V, Seidenspinner S, Zeitler P, et al.
Compound-heterozygous mutations in the plasminogen gene
predispose to the development of ligneous conjunctivitis. Blood
1999;93:34573466.
107. Ichinose A, Espling ES, Takamatsu J, et al. Two types of
abnormal genes for plasminogen in families with a
predisposition for thrombosis. Proc Natl Acad Sci U S A
1991;88:115119; and correction Proc Natl Acad Sci U S A
1991;88:2967.
108. Schott D, Dempfle C-E, Beck P, et al. Therapy with a
purified plasminogen concentrate in an infant with ligneous
conjunctivitis and homozygous plasminogen deficiency. N Engl J
Med 1998;339:16791686.
109. Azuma H, Uno Y, Shigekiyo T, et al. Congenital
plasminogen deficiency caused by a Ser 572 to Pro mutation.
Blood 1993;82:475480.
110. Takeda-Shitaka M, Umeyama H. Elucidation of the cause
for reduced activity of abnormal human plasmin containing an
Ala 55 Thr mutation: importance of highly conserved Ala 55 in
serine proteases. FEBS Lett 1998; 425:448452.
111. Mima N, Azuma H, Shigekiyo T, et al. A novel missense
mutation in two families with congenital plasminogen
deficiencyidentification of an Ala(675) to Thr(675)
substitution. Thromb Haemost 1996;75:96100.
112. Tsutsumi S, Saito T, Sakata T, et al. Genetic diagnosis of

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

80 de 138

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

dysplasminogenemia: detection of an Ala601-Thr mutation in


118 out of 125 families and identification of a new Asp676-Asn
mutation. Thromb Haemost 1996;76: 135138.
113. Yamaguchi M, Sugiyama S, Noda H, et al. A novel
missense mutation D676N in the plasminogen gene causes loss
of functional activity. Hum Hered 1997;47:234236.
114. Higuchi Y, Furihata K, Ueno I, et al. Plasminogen
Kanagawa-I, a novel missense mutation, is caused by the amino
acid substitution G732R. Br J Haematol 1998;103:867870.
115. Tait RC, Walker ID, Conkie JA, et al. Isolated familial
plasminogen deficiency may not be a risk factor for thrombosis.
Thromb Haemost 1996; 76:10041008.
116. Demarmels Biasiutti F, Sulzer I, Stucki B, et al. Is
plasminogen deficiency a thrombotic risk factor? A study of 23
thrombophilic patients and their family members. Thromb
Haemost 1998;80:167170.
117. Sartori TM, Saggiorato G, Pellati D, et al. Contraceptive
pills induce an improvement in congenital hypoplasminogenemia
in two unrelated patients with ligneous conjunctivitis. Thromb
Haemost 2003;90:8691.
118. Tefs K, Ziegler M, Georgieva M, et al. A nucleic acid
exchange in Intron F (Intron F-14T > G) in the human
plasminogen gene is only a common polymorphism and not a
true mutation. Thromb Haemost 2004;91:830831.
P.357
119. Drew AF, Kaufman AH, Kombrinck KW, et al. Ligneous
conjunctivitis in plasminogen-deficient mice. Blood
1998;91:16161624.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

81 de 138

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

120. Ploplis VA, Carmeliet P, Vazizadeh S, et al. Effects of


disruption of the plasminogen gene on thrombosis, growth, and
health in mice. Circulation 1995;92:25852593.
121. Bugge TH, Flick MJ, Daugherty CC, et al. Plasminogen
deficiency causes severe thrombosis but is compatible with
development and reproduction. Genes Dev 1995;9:794807.
122. Ny A, Leonardsson G, Hgglund A-C, et al. Ovulation in
plasminogen-deficient mice. Endocrinology
1999;140:50305035.
123. Rmer J, Bugge TH, Pyke C, et al. Impaired wound healing
in mice with a disrupted plasminogen gene. Nat Med
1996;2:287292.
124. Bezerra JA, Bugge TH, Melin-Aldana H, et al. Plasminogen
deficiency leads to impaired remodeling after a toxic injury to
the liver. Proc Natl Acad Sci U S A 1999;96:1514315148.
125. Ploplis VA, French EL, Carmeliet P, et al. Plasminogen
deficiency differentially affects recruitment of inflammatory cell
populations in mice. Blood 1998;91:20052009.
126. Carmeliet P, Moons L, Ploplis V, et al. Impaired arterial
neointima formation in mice with disruption of the plasminogen
gene. J Clin Invest 1997; 99:200208.
127. Xiao Q, Danton MJ, Witte DP, et al. Plasminogen deficiency
accelerates vessel wall disease in mice predisposed to
atherosclerosis. Proc Natl Acad Sci U S A
1997;94:1033510340.
128. Bugge TH, Kombrinck KW, Flick MJ, et al. Loss of
fibrinogen rescues mice from the pleiotropic effects of

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

82 de 138

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

plasminogen deficiency. Cell 1996;87: 709719.


129. Palumbo JS, Talmage KE, Liu H, et al. Plasminogen
supports tumor growth through a fibrinogen-dependent
mechanism linked to vascular patency. Blood
2003;102:28192827.
130. Bannach FG, Gutierrez A, Fowler BJ, et al. Localization of
regulatory elements mediating constitutive and
cytokine-stimulated plasminogen gene expression. J Biol Chem
2002;277:3857938588.
131. Kida M, Wakabayashi S, Ichinose A. Expression and
induction by IL-6 of the normal and variant genes for human
plasminogen. Biochem Biophys Res Commun
1997;230:129132.
132. Meroni G, Buraggi G, Mantovani R, et al. Motifs resembling
hepatocyte nuclear factor 1 and activator protein 3 mediate the
tissue specificity of the human plasminogen gene. Eur J
Biochem 1996;236:373382.
133. Strickland S. Tissue plasminogen activator in nervous
system function and dysfunction. Thromb Haemost
2001;86:138143.
134. Rosenberg S. The urokinase-type plasminogen activator
system in cancer and other pathological conditions: introduction
and perspective. Curr Pharm Des 2003;9:4p.
135. Levin EG, Loskutoff DJ. Cultured bovine endothelial cells
produce both urokinase and tissue-type plasminogen activators.
J Cell Biol 1982;94: 631636.
136. Nicoloso G, Hauert J, Kruithof EKO, et al. Fibrinolysis in
normal subjectscomparison between plasminogen activator

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

83 de 138

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

inhibitor and other components of the fibrinolytic system.


Thromb Haemost 1988;59:299303.
137. Stalder M, Hauert J, Kruithof EKO, et al. Release of
vascular plasminogen activator (vPA) after venous stasis:
electrophoretic-zymographic analysis of free and complexed
vPA. Br J Haematol 1985;61:169176.
138. Booth NA, Walker E, Maughan R, et al. Plasminogen
activator in normal subjects after exercise and venous
occlusion: tPA circulates as complexes with C1 inhibitor and
PAI-1. Blood 1987;69:13541362.
139. Benham FJ, Spurr N, Povey S, et al. Assignment of
tissue-type plasminogen activator to chromosome 8 in man and
identification of a common restriction length polymorphism
within the gene. Mol Biol Med 1984;2: 251259.
140. Degen SJF, Raiput B, Reich E. The human tissue
plasminogen activator gene. J Biol Chem 1986;261:69726985.
141. Jrnvall H, Pohl G, Bergsdorf N, et al. Differential
proteolysis and evidence for a residue exchange in tissue
plasminogen activator suggest possible association between two
types of protein microheterogeneity. FEBS Lett
1983;156:4750.
142. Pennica D, Holmes WE, Kohr WJ, et al. Cloning and
expression of human tissue-type plasminogen activator cDNA in
E. coli. Nature 1983;301: 214221.
143. Van Zonneveld A-J, Veerman H, Pannekoek H. On the
interaction of the finger- and kringle 2-domain of tissue-type
plasminogen activator with fibrin. Inhibition of kringle-2 binding
to fibrin by -aminocaproic acid. J Biol Chem
1986;261:1421414218.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

84 de 138

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

144. Bnyai L, Vradi A, Patthy L. Common evolutionary origin


of the fibrin-binding structures of fibronectin and tissue-type
plasminogen activator. FEBS Lett 1983;163:3741.
145. de Munk GAW, Caspers MPM, Chang GTG, et al. Binding of
tissue-type plasminogen activator to lysine, lysine analogues,
and fibrin fragments. Biochemistry 1989;28:73187325.
146. Downing AK, Driscoll PC, Harvey TS, et al. Solution
structure of the fibrin binding finger domain of tissue-type
plasminogen activator determined by nuclear magnetic
resonance. J Mol Biol 1992;225: 821833.
147. Smith BO, Downing AK, Dudgeon TJ, et al. Secondary
structure of fibronectin type 1 and epidermal growth factor
modules from tissue-type plasminogen activator by nuclear
magnetic resonance. Biochemistry 1994; 33:24222429.
148. Byeon I-JL, Kelley RF, Mulkerrin MG, et al. Ligand binding
to the tissue-type plasminogen activator kringle 2 domain:
structural characterization by

H-NMR. Biochemistry

1995;34:27392750.
149. Kelley RF, DeVos AM, Cleary S. Thermodynamics of ligand
binding and denaturation for His64 mutants of tissue
plasminogen activator kringle-2 domain. Proteins: Struct Funct
Genet 1991;11:3544.
150. de Vos AM, Ultsch MH, Kelley RF, et al. Crystal structure of
the kringle 2 domain of tissue plasminogen activator at 2.4-
resolution. Biochemistry 1992;31:270279.
151. Collen D, Lijnen HR, Bulens F, et al. Biochemical and
functional characterization of human tissue-type plasminogen
activator variants with mutagenized kringle domains. J Biol
Chem 1990;265:1218412191.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

85 de 138

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

152. Chang Y, Nilsen SL, Castellino FJ. Functional and structural


consequences of aromatic residue substitutions within the
kringle-2 domain of tissue-type plasminogen activator. J Pept
Res 1999;53:656664.
153. de Serrano VS, Sehl LC, Castellino FJ. Direct identification
of lysine-33 as the principal cationic center of the -amino acid
binding site of the recombinant kringle 2 domain of tissue-type
plasminogen activator. Arch Biochem Biophys
1992;292:206212.
154. Lamba D, Bauer M, Huber R, et al. The 2.3 crystal
structure of the catalytic domain of recombinant two-chain
human tissue-type plasminogen activator. J Mol Biol
1996;258:117135.
155. Rijken DC, Hoylaerts M, Collen D. Fibrinolytic properties of
one-chain and two-chain human extrinsic (tissue-type)
plasminogen activator. J Biol Chem 1982;257:29202925.
156. Tachias K, Madison EL. Converting tissue-type
plasminogen activator into a zymogen. J Biol Chem
1996;271:2874928752.
157. Camiolo SM, Thorsen S, Astrup T. Fibrinogenolysis and
fibrinolysis with tissue plasminogen activator, urokinase,
streptokinase-activated human globulin, and plasmin. Proc Soc
Exp Biol Med 1971;138:277280.
158. Thorsen S, Glas-Greenwalt P, Astrup T. Differences in the
binding to fibrin of urokinase and tissue plasminogen activator.
Thromb Diath Haemorrh 1972;28:6574.
159. Higgins D, Vehar GA. Interaction of one-chain and
two-chain tissue plasminogen activator with intact and
plasmin-degraded fibrin. Biochemistry 1987;26:77867791.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

86 de 138

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

160. Larsen GR, Henson H, Blue Y. Variants of human


tissue-type plasminogen activator. Fibrin binding, fibrinolytic,
and fibrinogenolytic characterization of genetic variants lacking
the fibronectin finger-like and/or the epidermal growth factor
domains. J Biol Chem 1988;263:10231029.
161. Rnby M, Bergsdorf N, Nilsson T. Enzymatic properties of
the one- and two-chain form of tissue plasminogen activator.
Thromb Res 1982;27: 175183.
162. Geppert AG, Binder BR. Allosteric regulation of
tPA-mediated plasminogen activation by a modifier mechanism:
evidence for a binding site for plasminogen on the tPA A-chain.
Arch Biochem Biophys 1992;297: 205212.
163. Rnby M. Studies on the kinetics of plasminogen activation
by tissue plasminogen activator. Biochim Biophys Acta
1982;704:461469.
164. Hoylaerts M, Rijken DC, Lijnen HR, et al. Kinetics of the
activation of plasminogen by human tissue plasminogen
activator. Role of fibrin. J Biol Chem 1982;257:29122919.
165. Norrman B, Walln P, Rnby M. Fibrinolysis mediated by
tissue plasminogen activator. Disclosure of a kinetic transition.
Eur J Biochem 1985; 149:193200.
166. Rijken DC, Wijngaards G, Zaal De Jong M, et al.
Purification and partial characterization of plasminogen
activator from human uterine tissue. Biochim Biophys Acta
1979;580:140153.
167. De Vries C, Veerman H, Koornneef E, et al. Tissue-type
plasminogen activator and its substrate Glu-plasminogen share
common binding sites in limited plasmin-digested fibrin. J Biol
Chem 1990;265:1354713552.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

87 de 138

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

168. Medved L, Nieuwenhuizen W. Molecular mechanisms of


initiation of fibrinolysis by fibrin. Thromb Haemost
2003;89:409419.
169. Yonekawa O, Voskuilen M, Nieuwenhuizen W. Localization
in the fibrinogen -chain of a new site that is involved in the
acceleration of the tissue-type plasminogen activator-catalysed
activation of plasminogen. Biochem J 1992;283:187191.
170. Schielen WJG, Adams HPHM, van Leuven K, et al. The
sequence -(312-324) is a fibrin-specific epitope. Blood
1991;77:21692173.
171. Tsurupa G, Medved L. Identification and characterization
of novel tPA- and plasminogen-binding sites within fibrin(ogen)
C-domain. Biochemistry 2001;40:801808.
172. Mosesson MW, Siebenlist KR, Voskuilen M, et al.
Evaluation of the factors contributing to fibrin-dependent
plasminogen activation. Thromb Haemost 1998;79:796801.
173. Madison EL, Goldsmith EJ, Gerard RD, et al.
Serpin-resistant mutants of human tissue-type plasminogen
activator. Nature 1989;339:721724.
174. Madison EL, Goldsmith EJ, Gerard RD, et al. Amino acid
residue that affect interaction of tissue-type plasminogen
activator with plasminogen activator inhibitor 1. Proc Natl Acad
Sci U S A 1990;87:35303533.
175. Paoni NF, Keyt BA, Refino CJ, et al. A slow clearing,
fibrin-specific, PAI-1 resistant variant of tPA (T103N, KHRR
296-299 AAAA). Thromb Haemost 1993;70:307312.
176. Stewart RJ, Fredenburgh JC, Leslie BA, et al. Identification

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

88 de 138

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

of the mechanism responsible for the increased fibrin specificity


of TNK-tissue plasminogen activator relative to tissue
plasminogen activator. J Biol Chem 2000;275:1011210120.
P.358
177. Hajjar KA, Krishnan S. Annexin II: a mediator of the
plasmin/plasminogen activator system. Trends Cardiovasc Med
1999;9:128138.
178. Dudani AK, Ganz PR. Endothelial cell surface actin serves
as a binding site for plasminogen, tissue plasminogen activator
and lipoprotein(a). Br J Haematol 1996;95:168178.
179. Bohm T, Geiger M, Binder BR. Isolation and
characterization of tissue-type plasminogen activator-binding
proteoglycans from human umbilical vein endothelial cells.
Arterioscler Thromb Vasc Biol 1996;16: 665672.
180. Hembrough TA, Kralovich KR, Li L, et al. Cytokeratin 8
released by breast carcinoma cells in vitro binds plasminogen
and tissue-type plasminogen activator and promotes
plasminogen activation. Biochem J 1996;317: 763769.
181. Beebe DP, Wood LL, Moos M. Characterization of tissue
plasminogen activator binding proteins isolated from endothelial
cells and other cell types. Thromb Res 1990;59:339350.
182. Menell JS, Cesarman GM, Jacovina AT, et al. Annexin II
and bleeding in acute promyelocytic leukemia. N Engl J Med
1999;340:9941004.
183. Bennett B, Booth NA, Croll A, et al. The bleeding disorder
in acute promyelocytic leukaemia: fibrinolysis due to uPA rather
than defibrination. Br J Haematol 1989;71:511517.
184. Redlitz A, Plow EF. Receptors for plasminogen and tPA: an

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

89 de 138

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

update. Baillieres Clin Haematol 1995;8:313327.


185. Fukao H, Matsuo O. Analysis of tissue-type plasminogen
activator receptor (tPAR) in human endothelial cells. Semin
Thromb Hemost 1998;24: 269273.
186. Werner F, Razzaq TM, Ellis V. Tissue plasminogen activator
binds to human vascular smooth muscle cells by a novel
mechanism. Evidence for a reciprocal linkage between inhibition
of catalytic activity and cellular binding. J Biol Chem
1999;274:2155521561.
187. Razzaq TM, Bass R, Vines DJ, et al. Functional regulation
of tissue plasminogen activator on the surface of vascular
smooth muscle cells by the type-II transmembrane protein p63
(CKAP4). J Biol Chem 2003;278: 4267942685.
188. Parekh RB, Dwek RA, Thomas JR, et al. Cell-type-specific
and site-specific N-glycosylation of type I and type II human
tissue plasminogen activator. Biochemistry
1989;28:76447662.
189. Dwek RA. Glycobiology: Towards understanding the
function of sugars. Biochem Soc Trans 1995;23:125.
190. Harris RJ, Leonard CK, Guzzetta AW, et al. Tissue
plasminogen activator has an O-linked fucose attached to
threonine-61 in the epidermal growth factor domain.
Biochemistry 1991;30:23112314.
191. Hajjar KA, Reynolds CM. -Fucose-mediated binding and
degradation of tissue-type plasminogen activator by HepG2
cells. J Clin Invest 1994;93: 703710.
192. Ridker PM, Baker MT, Hennekens CH, et al. Alu-repeat
polymorphism in the gene coding for tissue-type plasminogen

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

90 de 138

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

activator (tPA) and risks of myocardial infarction among


middle-aged men. Arterioscler Thromb Vasc Biol
1997;17:16871690.
193. van der Bom JG, de Knijff P, Haverkate F et al. The
Rotterdam Study. Tissue plasminogen activator and risk of
myocardial infarction. Circulation 1997;95:26232627.
194. Jern C, Ladenvall P, Wall U, et al. Gene polymorphism of
t-PA is associated with forearm vascular release rate of t-PA.
Arterioscler Thromb Vasc Biol 1999;19:454459.
195. Tjarnlund-Wolf A, Medcalf RL, Jern C. The t-PA -7,351C >
T enhancer polymorphism decreases Sp1 and Sp3 protein
binding affinity and transcriptional responsiveness to retinoic
acid. Blood 2005;105:10601067.
196. Arts J, Herr I, Lansink M, et al. Cell-type specific
DNA-protein interactions at the tissue-type plasminogen
activator promoter in human endothelial and Hela cells in vivo
and in vitro. Nucleic Acids Res 1997;25:311317.
197. Levin EG, Santell L. Stimulation and desensitization of
tissue plasminogen activator release from human endothelial
cells. J Biol Chem 1988;263: 93609365.
198. Levin EG, Marotti KR, Santell L. Protein kinase C and the
stimulation of tissue plasminogen activator release from human
endothelial cells: dependence on the elevation of messenger
RNA. J Biol Chem 1989;264: 1603016036.
199. Bulens F, Merchiers P, Ibaez-Tallon I, et al. Identification
of a multihormone responsive enhancer far upstream from the
human tissue-type plasminogen activator gene. J Biol Chem
1997;272:663671.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

91 de 138

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

200. Merchiers P, Bulens F, Stockmans I, et al.


1,25-Dihydroxyvitamin D 3 induction of the tissue-type
plasminogen activator gene is mediated through its
multihormone-responsive enhancer. FEBS Lett
1999;460:289296.
201. Kooistra T, Opdenberg JP, Toet K, et al. Stimulation of
tissue-type plasminogen activator synthesis by retinoids in
cultured human endothelial cells and rat tissues in vivo. Thromb
Haemost 1991;65:565572.
202. Medh RD, Santell L, Levin EG. Stimulation of tissue
plasminogen activator production by retinoic acid: synergistic
effect on protein kinase C-mediated activation. Blood
1992;80:981987.
203. Levin EG, del Zoppo GJ. Localization of tissue plasminogen
activator in the endothelium of a limited number of vessels. Am
J Pathol 1994;144: 855861.
204. Levin EG, Santell L, Osborn KG. The expression of tissue
plasminogen activator in vivo: a function defined by vessel size
and anatomic location. J Cell Sci 1997;110:139148.
205. Pham NL, Franzen A, Levin EG. NF1 regulatory element
functions as a repressor of tissue plasminogen activator
expression. Arterioscler Thromb Vasc Biol 2004;24:982987.
206. Costa M, Shen Y, Maurer F, et al. Transcriptional
regulation of the tissue-type plasminogen-activator gene in
human endothelial cells: identification of nuclear factors that
recognise functional elements in the tissue-type
plasminogen-activator gene promoter. Eur J Biochem
1998;258:123131.
207. Medcalf RL, Regg M, Schleuning WD. A DNA motif related

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

92 de 138

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

to the cAMP-responsive element and an exon-located activator


protein-2 binding site in the human tissue-type plasminogen
activator gene promoter cooperate in basal expression and
convey activation by phorbol ester and cAMP. J Biol Chem
1990;265:1461814626.
208. Fujiwara J, Kimura T, Ayusawa D, et al. A novel regulatory
sequence affecting the constitutive expression of tissue
plasminogen activator (tPA) gene in human melanoma (Bowes)
cells. J Biol Chem 1994;269:1855818562.
209. Bulens F, Ibaez-Tallon I, van Acker P, et al. Retinoic acid
induction of human tissue-type plasminogen activator gene
expression via a direct repeat element (DR5) located at -7
kilobases. J Biol Chem 1995;270: 71677175.
210. Keber D, Blinc A, Fettich J. Increase of tissue plasminogen
activator in limbs during venous occlusion: a simple
haemodynamic model. Thromb Haemost 1990;64:433437.
211. Juhan-Vague I, Valadier J, Alessi MC, et al. Deficient tPA
release and elevated PA inhibitor levels in patients with
spontaneous or recurrent deep venous thrombosis. Thromb
Haemost 1987;57:6772.
212. Nguyen G, Horellou MH, Kruithof EKO, et al. Residual
plasminogen activator inhibitor activity after venous stasis as
criterion for hypofibrinolysis: a study in 83 patients with
confirmed deep vein thrombosis. Blood 1988;72:601605.
213. Jeanneau C, Bachouchi NO, Gorin I, et al. Absence of
functional activity of tissue plasminogen activator in patients
with severe forms of von Willebrand's disease. Br J Haematol
1987;56:7988.
214. Knop M, Aareskjold E, Bode G, et al. Rab3D and annexin

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

93 de 138

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

A2 play a role in regulated secretion of VWF, but not tPA, from


endothelial cells. EMBO J 2004;23:29822992.
215. Keber D, Stegnar M, Kluft C. Different tissue plasminogen
activator release in the arm and leg during venous occlusion is
equalized after DDAVP infusion. Thromb Haemost
1990;63:7275.
216. Waugh JM, Kattash M, Li J, et al. Gene therapy to promote
thromboresistance: local overexpression of tissue plasminogen
activator to prevent arterial thrombosis in an in vivo rabbit
model. Proc Natl Acad Sci U S A 1999;96:10651070.
217. Booth NA, Bennett B, Wijngaards G, et al. A new life-long
hemorrhagic disorder due to excess plasminogen activator.
Blood 1983;61:267275.
218. Aznar J, Estells A, Vila V, et al. Inherited fibrinolytic
disorder due to an enhanced plasminogen activator level.
Thromb Haemost 1984;52:196200.
219. Emeis JJ, van den Eijnden-Schrauwen Y, Van den Hoogen
CM, et al. An endothelial storage granule for tissue-type
plasminogen activator. J Cell Biol 1997;139:245256.
220. Parmer RJ, Mahata M, Mahata S, et al. Tissue plasminogen
activator (tPA) is targeted to the regulated secretory pathway.
Catecholamine storage vesicles as a reservoir for the rapid
release of tPA. J Biol Chem 1997; 272:19761982.
221. Rosnoblet C, Vischer UM, Gerard RD, et al. Storage of
tissue-type plasminogen activator in Weibel-Palade bodies of
human endothelial cells. Arterioscler Thromb Vasc Biol
1999;19:17961803.
222. Santell L, Marotti KR, Levin EG. Targeting of tissue

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

94 de 138

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

plasminogen activator into the regulated secretory pathway of


neuroendocrine cells. Brain Res 1999;816:258265.
223. Padr T, Van den Hoogen CM, Emeis JJ. Distribution of
tissue-type plasminogen activator (activity and antigen) in rat
tissues. Blood Coagul Fibrinolysis 1990;1:601608.
224. Tappy L, Hauert J, Bachmann F. Effects of hypoxia and
acidosis on vascular plasminogen activator release in the pig
ear perfusion system. Thromb Res 1984;33:117124.
225. Vaughan DE. The renin-angiotensin system and
fibrinolysis. Am J Cardiol 1997;79:1216.
226. van den Eijnden-Schrauen Y, Atsma DE, Lupu F, et al.
Involvement of calcium and G proteins in the acute release of
tissue-type plasminogen activator and von Willebrand factor
from cultured human endothelial cells. Arterioscler Thromb Vasc
Biol 1997;17:21772187.
227. Chandler WL, Levy WC, Stratton JR. The circulatory
regulation of TPA and UPA secretion, clearance, and inhibition
during exercise and during the infusion of isoproterenol and
phenylephrine. Circulation 1995;92:29842994.
228. Jern C, Selin L, Jern S. In vivo release of tissue-type
plasminogen activator across the human forearm during mental
stress. Thromb Haemost 1994; 72:285291.
229. Cash JD. Control mechanisms of activator release. In:
Davidson JF, Rowan RM, Samama MM et al., eds. Progress in
chemical fibrinolysis and thrombolysis, Vol. 3. New York: Raven
Press, 1978:6575.
230. Mannucci PM, Rota L. Plasminogen activator response after
DDAVP: a clinico-pharmacological study. Thromb Res

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

95 de 138

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

1980;20:6976.
231. Brommer EJP, Barrett-Bergshoeff MM, Allen RA, et al. The
use of desmopressin acetate (DDAVP) as a test of fibrinolytic
capacity of patientsanalysis of responders and
non-responders. Thromb Haemost 1982;48: 156161.
P.359
232. Wall U, Jern S, Tengborn L, et al. Evidence of a local
mechanism for desmopressin-induced tissue-type plasminogen
activator release in human forearm. Blood 1998;91:529537.
233. Stein CM, Brown N, Vaughan DE, et al. Regulation of local
tissue-type plasminogen activator release by
endothelium-dependent and endothelium-independent agonists
in human vasculature. J Am Coll Cardiol 1998; 32:117122.
234. Dell'Omo G, Ferrini L, Morale M, et al.
Acetylcholine-mediated vasodilatation and tissue-type
plasminogen activator release in normal and hypertensive men.
Angiology 1999;50:273282.
235. Brown NJ, Gainer JV, Stein CM, et al. Bradykinin
stimulates tissue plasminogen activator release in human
vasculature. Hypertension 1999;33: 14311435.
236. Newby DE, Wright RA, Ludlam CA, et al. An in vivo model
for the assessment of acute fibrinolytic capacity of the
endothelium. Thromb Haemost 1997;78:12421248.
237. Hrafnkelsdottir T, Gudnason T, Wall U, et al. Regulation of
local availability of active tissue-type plasminogen activator in
vivo in man. J Thromb Haemost 2004;2:19601968.
238. Tanswell P, Seifried E, Su PCAF, et al. Pharmacokinetics
and systemic effects of tissue-type plasminogen activator in

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

96 de 138

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

normal subjects. Clin Pharmacol Ther 1989;46:155162.


239. Huber K, Kirchheimer JC, Korninger C, et al. Hepatic
synthesis and clearance of components of the fibrinolytic
system in healthy volunteers and in patients with different
stages of liver cirrhosis. Thromb Res 1991;62: 491500.
240. Wing LR, Hawksworth GM, Bennett B, et al. Clearance of
t-PA, PAI-1, and t-PA-PAI-1 complex in an isolated perfused rat
liver system. J Lab Clin Med 1991;117:109114.
241. Chandler WL, Alessi MC, Aillaud MF, et al. Clearance of
tissue plasminogen activator (TPA) and TPA/plasminogen
activator inhibitor type 1 (PAI-1) complex: relationship to
elevated TPA antigen in patients with high PAI-1 activity levels.
Circulation 1997;96:761768.
242. Bounameaux H, Stassen JM, Seghers C, et al. Influence of
fibrin and liver blood flow on the turnover and the systemic
fibrinogenolytic effects of recombinant human tissue-type
plasminogen activator in rabbits. Blood 1986;67:14931497.
243. de Boer A, Kluft C, Kroon JM, et al. Liver blood flow as a
major determinant of the clearance of recombinant human
tissue-type plasminogen activator. Thromb Haemost
1992;67:8387.
244. Van Griensven JMT, Huisman LGM, Stuurman T, et al.
Effects of increased liver blood flow on the kinetics and
dynamics of recombinant tissue-type plasminogen activator.
Clin Pharmacol Ther 1996;60:504511.
245. Fletcher AP, Biederman O, Moore D, et al. Abnormal
plasminogen-plasmin system activity (fibrinolysis) in patients
with hepatic cirrhosis: its cause and consequences. J Clin Invest
1964;43:681695.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

97 de 138

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

246. Otter M, Barrett-Bergshoeff MM, Rijken DC. Binding of


tissue-type plasminogen activator by the mannose receptor. J
Biol Chem 1991;266: 1393113935.
247. Biessen EAL, van Teijlingen M, Vietsch H, et al.
Antagonists of the mannose receptor and the LDL
receptor-related protein dramatically delay the clearance of
tissue plasminogen activator. Circulation 1997;95:4652.
248. Noorman F, Barrett-Bergshoeff MM, Rijken DC. Role of
carbohydrate and protein in the binding of tissue-type
plasminogen activator to the human mannose receptor. Eur J
Biochem 1998;251:107113.
249. Lansink M, Jong M, Bijsterbosch M, et al. Increased
clearance explains lower plasma levels of tissue-type
plasminogen activator by estradiol: evidence for potently
enhanced mannose receptor expression in mice. Blood
1999;94:13301336.
250. Lehrman MA, Hill RL. The binding of fucose-containing
glycoproteins by hepatic lectins. J Biol Chem
1986;261:74197474.
251. Strickland DK, Ranganathan S. Diverse role of LDL
receptor-related protein in the clearance of proteases and in
signaling. J Thromb Haemost 2003;1:16631670.
252. Narita M, Bu G, Herz J, et al. Two receptor systems are
involved in the plasma clearance of tissue-type plasminogen
activator (tPA) in vivo. J Clin Invest 1995;96:11641168.
253. Camani C, Bachmann F, Kruithof EKO. The role of
plasminogen activator inhibitor type 1 in the clearance of
tissue-type plasminogen activator by rat hepatoma cells. J Biol
Chem 1994;269:57705775.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

98 de 138

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

254. Moestrup SK, Holtet TL, Etzerodt M, et al.


2 -Macroglobulin complexes, plasminogen activator inhibitor
type-1-plasminogen activator complexes, and
receptor-associated protein bind to a region of the
2 macroglobulin receptor containing a cluster of eight
complement-type repeats. J Biol Chem 1993;268:1369113696.
255. Willnow TE, Orth K, Herz J. Molecular dissection of ligand
binding sites on the low density lipoprotein receptor-related
protein. J Biol Chem 1994; 269:1582715832.
256. Horn IR, Van den Berg BMM, van der Meijden PZ, et al.
Molecular analysis of ligand binding to the second cluster of
complement-type repeats of the low density lipoprotein
receptor-related protein. Evidence for an allosteric component
in receptor-associated protein-mediated inhibition of ligand
binding. J Biol Chem 1997;272:1360813613.
257. Herz J, Goldstein JL, Strickland DK, et al. 39-kDa protein
modulates binding of ligands to low density lipoprotein
receptor-related protein/ 2 macroglobulin receptor. J Biol Chem
1991;266:2123221238.
258. Andreasen PA, Sottrup-Jensen L, Kjller L, et al.
Receptor-mediated endocytosis of plasminogen activators and
activator/inhibitor complexes. FEBS Lett 1994;338:239245.
259. Moestrup SK. The 2 macroglobulin receptor and epithelial
glycoprotein-330: two giant receptors mediating endocytosis of
multiple ligands. Biochim Biophys Acta 1994;1197:197213.
260. Kasza A, Petersen HH, Heegaard CW, et al. Specificity of
serine proteinase/serpin complex binding to very-low-density
lipoprotein receptor and 2 macroglobulin
receptor/low-density-lipoprotein-receptor-related protein. Eur J
Biochem 1997;248:270281.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

99 de 138

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

261. Kounnas MZ, Henkin J, Argraves WS, et al. Low density


lipoprotein receptor-related portion/ 2 macroglobulin receptor
mediates cellular uptake of pro-urokinase. J Biol Chem
1993;268:2186221867.
262. Nykjaer A, Kjller L, Cohen RL, et al. Regions involved in
binding of urokinase-type-1 inhibitor complex and
pro-urokinase to the endocytic 2 macroglobulin receptor/low
density lipoprotein receptor-related protein. Evidence that the
urokinase receptor protects pro-urokinase against binding to
the endocytic receptor. J Biol Chem 1994;269:2566825676.
263. Camani C, Gavin O, Kruithof EKO. Cellular degradation of
free and inhibitor-bound tissue-type plasminogen activator.
Requirement for a co- receptor? Thromb Haemost
2000;83:290296.
264. Carmeliet P, Schoonjans L, Kieckens L, et al. Physiological
consequences of loss of plasminogen activator gene function in
mice. Nature 1994;368: 419424.
265. Carmeliet P, Collen D. Targeted gene manipulation and
transfer of the plasminogen and coagulation systems in mice.
Fibrinolysis 1996;10:195213.
266. Leonardsson G, Peng X-R, Liu K, et al. Ovulation efficiency
is reduced in mice that lack plasminogen activator gene
function: functional redundancy among physiological
plasminogen activators. Proc Natl Acad Sci U S A
1995;92:1244612450.
267. Seeds NW, Basham ME, Haffke SP. Neuronal migration is
retarded in mice lacking the tissue plasminogen activator gene.
Proc Natl Acad Sci U S A 1999;96:1411814123.
268. Baranes D, Lederfein D, Huang Y-Y, et al. Tissue

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

100 de 138

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

plasminogen activator contributes to the late phase of LTP and


to synaptic growth in the hippocampal mossy fiber pathway.
Neuron 1998;21:813825.
269. Husain SS, Gurewich V, Lipinski B. Purification and partial
characterization of a single-chain molecular weight form of
urokinase from human urine. Arch Biochem Biophys
1983;220:3138.
270. Larsson L, Skriver L, Nielsen LS, et al. Distribution of
urokinase-type plasminogen activator immunoreactivity in the
mouse. J Cell Biol 1984; 98:894903.
271. Grau E, Moroz LA. Fibrinolytic activity of normal human
blood monocytes. Thromb Res 1989;53:145162.
272. Manchanda N, Schwartz BS. Lipopolysaccharide-induced
modulation of human monocyte urokinase production and
activity. J Immunol 1990; 145:41744180.
273. Van Hinsbergh VWM, Kooistra T, van den Berg EA, et al.
Tumor necrosis factor increases the production of plasminogen
activator inhibitor in human endothelial cells in vitro and in rats
in vivo. Blood 1988;72: 14671473.
274. Camoin L, Pannell R, Anfosso F, et al. Evidence for the
expression of urokinase-type plasminogen activator by human
venous endothelial cells in vivo. Thromb Haemost
1998;80:961967.
275. Dan K, Andreasen PA, Grndahl-Hansen J, et al.
Plasminogen activators, tissue degradation, and cancer. Adv
Cancer Res 1985;44:139266.
276. Duffy MJ. The urokinase plasminogen activator system:
role in malignancy. Curr Pharm Des 2004;10:3949.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

101 de 138

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

277. Steffens GJ, Gnzler WA, tting F, et al. The complete


amino acid sequence of low molecular mass urokinase from
human urine. Hoppe Seylers Z Physiol Chem
1982;363:10431058.
278. Holmes WE, Pennica D, Blaber M, et al. Cloning and
expression of the gene for pro-urokinase in Escherichia coli.
Biotechnology 1985;3:923929.
279. Riccio A, Grimaldi G, Verde P, et al. The human urokinase
plasminogen activator gene and its promoter. Nucleic Acids Res
1985;13:27532771.
280. Triputti P, Blasi F, Verde P, et al. Human urokinase gene is
located on the long arm of chromosome 10. Proc Natl Acad Sci
U S A 1985;82: 44484452.
281. Bdeir K, Kuo A, Sachais BS, et al. The kringle stabilizes
urokinase binding to the urokinase receptor. Blood
2003;102:36003608.
282. Buko AM, Kentzer EJ, Petros A, et al. Characterization of a
posttranslational fucosylation in the growth factor domain of
urinary plasminogen activator. Proc Natl Acad Sci U S A
1991;88:39923996.
283. Rabbani SA, Mazar AP, Bernier SM, et al. Structural
requirements for the growth factor activity of the
amino-terminal domain of urokinase. J Biol Chem
1992;267:1415114156.
284. Franco P, Iaccarino C, Chiaradonna F, et al.
Phosphorylation of human pro-urokinase on Ser138/303 impairs
its receptor-dependent ability to promote myelomonocytic
adherence and motility. J Cell Biol 1997;137: 779791.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

102 de 138

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

285. Lijnen HR, Van Hoef B, Collen D. Activation with plasmin of


two-chain urokinase-type plasminogen activator derived from
single-chain urokinase-type plasminogen activator by treatment
with thrombin. Eur J Biochem 1987;169:359364.
286. Bernik MB, Oller EP. Increased plasminogen activator
(urokinase) in tissue culture after fibrin deposition. J Clin
Invest 1973;42:823834.
P.360
287. Ichinose A, Fujikawa K, Suyama T. The activation of
pro-urokinase by plasma kallikrein and its inactivation by
thrombin. J Biol Chem 1986; 261:34863489.
288. List K, Jensen ON, Bugge TH, et al.
Plasminogen-independent initiation of the pro-urokinase
activation cascade in vivo. Activation of pro-urokinase by
glandular kallikrein (mGK-6) in plasminogen-deficient mice.
Biochemistry 2000;39:508515.
289. Gurewich V, Pannell R. Inactivation of single-chain
urokinase (pro-urokinase) by thrombin and thrombin-like
enzymes: relevance of the findings to the interpretation of
fibrin-binding experiments. Blood 1987;69:769772.
290. Braat EAM, Levi M, Bos R, et al. Inactivation of
single-chain urokinase-type plasminogen activator by thrombin
in human subjects. J Lab Clin Med 1999;134:161167.
291. Nauland U, Rijken DC. Activation of thrombin-inactivated
single-chain urokinase-type plasminogen activator by dipeptidyl
peptidase I (cathepsin C). Eur J Biochem 1994;223:497501.
292. Schmitt M, Kanayama N, Henschen A, et al. Elastase
released from human granulocytes stimulated with
N-formyl-chemotactic peptide prevents activation of tumor cell

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

103 de 138

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

prourokinase (pro-uPA). FEBS Lett 1989; 255:8388.


293. Gnzler WA, Steffens GJ, tting F, et al. Structural
relationship between human high and low molecular mass
urokinase. Hoppe Seylers Z Physiol Chem 1982;363:133141.
294. Learmonth MP, Li W, Namiranian S, et al. Modulation of
the cell binding property of single chain urokinase-type
plasminogen activator by neutrophil cathepsin G. Fibrinolysis
1992;6(Suppl. 4):113116.
295. Marcotte PA, Henkin J, Credo RB, et al. A-chain isozymes
of recombinant and natural urokinases: preparation,
characterization, and their biochemical and fibrinolytic
properties. Fibrinolysis 1992;6:6978.
296. Li X, Bokman AM, Llins M, et al. Solution structure of the
kringle domain from urokinase-type plasminogen activator. J
Mol Biol 1994;235: 15481559.
297. Hansen AP, Petros AM, Meadows RP, et al. Solution
structure of the amino-terminal fragment of urokinase-type
plasminogen activator. Biochemistry 1994;33:48474864.
298. Spraggon G, Phillips C, Nowak UK, et al. The crystal
structure of the catalytic domain of human urokinase-type
plasminogen activator. Structure 1995;3:681691.
299. Sun Z, Liu BF, Chen Y, et al. Analysis of the forces which
stabilize the active conformation of urokinase-type plasminogen
activator. Biochemistry 1998;37:29352940.
300. Zeslawska E, Schweinitz A, Karcher A, et al. Crystals of
the urokinase type plasminogen activator variant (c)-uPA in
complex with small molecule inhibitors open the way towards
structure-based drug design. J Mol Biol 2000;301:465475.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

104 de 138

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

301. Schweinitz A, Steinmetzer T, Banke IJ, et al. Design of


novel and selective inhibitors of urokinase-type plasminogen
activator with improved pharmacokinetic properties for use as
antimetastatic agents. J Biol Chem 2004;279:3361333622.
302. Petersen LC. Kinetics of reciprocal
pro-urokinase/plasminogen activation. Stimulation by a
template formed by the urokinase receptor bound to
poly(D-lysine). Eur J Biochem 1997;245:316323.
303. Lijnen HR, Van Hoef B, Nelles L, et al. Plasminogen
activation with single-chain urokinase-type plasminogen
activator scuPA. Studies with active site mutagenized
plasminogen (Ser 158 Glu) and plasmin-resistant scuPA
(Lys 158 Glu). J Biol Chem 1990;265:52325236.
304. Pannell R, Gurewich V. Activation of plasminogen by
single-chain urokinase or by two-chain urokinasea
demonstration that single-chain urokinase has a low catalytic
activity (pro-urokinase). Blood 1987;69:2226.
305. Lijnen HR, Van Hoef B, De Cock F, et al. The mechanism of
plasminogen activation and fibrin dissolution by single chain
urokinase-type plasminogen activator in a plasma milieu in
vitro. Blood 1989;73:18641872.
306. Gurewich V, Pannell R, Louie S, et al. Effective and
fibrin-specific clot lysis by a zymogen precursor form of
urokinase (pro-urokinase): a study in vitro and in two animal
species. J Clin Invest 1984;73:17311739.
307. Ellis V, Scully MF, Kakkar VV. Plasminogen activation by
single-chain urokinase in functional isolation. J Biol Chem
1987;262:1499815003.
308. Liu J-N, Tang W, Sun Z-Y, et al. A site-directed

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

105 de 138

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

mutagenesis of pro- urokinase which substantially reduces its


intrinsic activity. Biochemistry 1996;35:1407014076.
309. Sun Z, Jiang Y, Ma Z, et al. Identification of a flexible loop
region (297-313) of urokinase-type plasminogen activator,
which helps determine its catalytic activity. J Biol Chem
1997;272:2381823823.
310. Vassalli J, Baccino D, Belin D. A cellular binding site for
the M r 55,000 form of the human plasminogen activator,
urokinase. J Cell Biol 1985; 100:8692.
311. Brglum AD, Byskov A, Ragno P, et al. Assignment of the
urokinase-type plasminogen activator receptor gene (PLAUR) to
chromosome 19q13.1-q 13.2. Am J Hum Genet
1992;50:492497.
312. Casey JR, Petranka JG, Kottra J, et al. The structure of the
urokinase-type plasminogen activator receptor gene. Blood
1994;84:11511156.
313. Dang J, Boyd D, Wang H, et al. A region between and -141
and -61 bp containing a proximal AP-1 is essential for
constitutive expression of urokinase-type plasminogen activator
receptor. Eur J Biochem 1999;264:9299.
314. Shetty S, Muniyappa H, Haldy PK, et al. Regulation of
urokinase receptor expression by phosphoglycerate kinase. Am
J Respir Cell Mol Biol 2004; 31:100106.
315. Ploug M. Identification of specific sites involved in ligand
binding by photoaffinity labeling of the receptor for the
urokinase-type plasminogen activator. Residues located at
equivalent positions in uPAR domains I and III participate in the
assembly of a composite ligand-binding site. Biochemistry
1998;37:1649416505.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

106 de 138

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

316. Oda M, Shiraishi A, Hasegawa M. Analysis of the ternary


complex formation of human urokinase with the separated two
domains of its receptor. Eur J Biochem 1998;256:411418.
317. Grdsvoll H, Dan K, Ploug M. Mapping part of the
functional epitope for ligand binding on the receptor for
urokinase-type plasminogen activator by site-directed
mutagenesis. J Biol Chem 1999;274:3799538003.
318. Behrendt N, Rnne E, Dan K. The structure and function
of the urokinase receptor, a membrane protein governing
plasminogen activation on the cell surface. Biol Chem Hoppe
Seyler 1995;376:269279.
319. Ellis V. Functional analysis of the cellular receptor for
urokinase in plasminogen activation. Receptor binding has no
influence on the zymogenic nature of pro-urokinase. J Biol
Chem 1996;271:1477914784.
320. Webb DJ, Nguyen DH, Gonias SL. Extracellular
signal-regulated kinase functions in the urokinase
receptor-dependent pathway by which neutralization of low
density lipoprotein receptor-related protein promotes
fibrosarcoma cell migration and matrigel invasion. J Cell Sci
2000;113: 123134.
321. Wei Y, Yang X, Liu Q, et al. A role for caveolin and the
urokinase receptor in integrin-mediated adhesion and signaling.
J Cell Biol 1999;144:12851294.
322. Wienke D, MacFadyen JR, Isacke CM. Identification and
characterization of the endocytic transmembrane glycoprotein
Endo180 as a novel collagen receptor. Mol Biol Cell
2003;14:35923604.
323. Kjoller L, Engelholm LH, Hyer-Hansen M, et al.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

107 de 138

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

uPARAP/endo180 directs lysosomal delivery and degradation of


collagen IV. Exp Cell Res 2004;293:106116.
324. Nykjaer A, Conese M, Christensen EI, et al. Recycling of
the urokinase receptor upon internalization of the uPA:serpin
complexes. EMBO J 1997; 16:26102620.
325. Wun TC, Schleuning D, Reich E. Isolation and
characterization of urokinase from human plasma. J Biol Chem
1982;257:32763283.
326. Darras V, Thienpont M, Stump DC, et al. Measurement of
urokinase-type plasminogen activator (uPA) with an
enzyme-linked immunosorbent assay (ELISA) based on three
murine monoclonal antibodies. Thromb Haemost
1986;56:411414.
327. Andreotti F, Kluft C. Circadian variation of fibrinolytic
activity in blood. Chronobiol Int 1991;8:336351.
328. Levi M, ten Cate JW, Dooijewaard G, et al. DDAVP induces
systemic release of urokinase-type plasminogen activator.
Thromb Haemost 1989;62: 686689.
329. Dooijewaard G, de Boer A, Turion PNC, et al. Physical
exercise induces enhancement of urokinase-type plasminogen
activator (uPA) levels in plasma. Thromb Haemost
1991;65:8286.
330. Booth NA, Anderson JA, Bennett B. Plasminogen activators
in alcoholic cirrhosis: demonstration of increased tissue type
and urokinase type activator. J Clin Pathol 1984;37:772777.
331. van der Kaaden ME, Rijken DC, Van Berkel TJC, et al.
Plasma clearance of urokinase-type plasminogen activator.
Fibrinolysis and Proteolysis 1998; 12:251258.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

108 de 138

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

332. Ellis V, Scully MF, Kakkar VV. Plasminogen activation


initiated by single-chain urokinase-type plasminogen activator.
Potentiation by U937 monocytes. J Biol Chem
1989;264:21852188.
333. Manchanda N, Schwartz BS. Single chain urokinase.
Augmentation of enzymatic activity upon binding to monocytes.
J Biol Chem 1991;266: 1458014584.
334. McGuinness CL, Humphries J, Waltham M, et al.
Recruitment of labelled monocytes by experimental venous
thrombi. Thromb Haemost 2001;85: 10181024.
335. Grau E, Moroz LA. Fibrinolytic activity of normal human
blood monocytes. Thromb Res 1989;53:145162.
336. Mutch NJ, Moir E, Robbie LA, et al. Localization and
identification of thrombin and plasminogen activator activities
in model human thrombi by in situ zymography. Thromb
Haemost 2002;88:9961002.
337. Moir E, Booth NA, Bennett B, et al. Polymorphonuclear
leucocytes mediate endogenous thrombus lysis via a
u-PA-dependent mechanism. Br J Haematol 2001;113:7280.
338. Moir E, Robbie LA, Bennett B, et al. Polymorphonuclear
leucocytes have two opposing roles in fibrinolysis. Thromb
Haemost 2002;87: 10061010.
339. Pluskota E, Soloviev DA, Bdeir K, et al. Integrin M 2
orchestrates and accelerates plasminogen activation and
fibrinolysis by neutrophils. J Biol Chem 2004;279:1806318072.
340. Manchanda N, Schwartz BS. Interaction of single-chain
urokinase and plasminogen activator inhibitor type 1. J Biol

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

109 de 138

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

Chem 1995;270:2003220035.
341. Schwartz BS, Espaa F. Two distinct urokinase-serpin
interactions regulate the initiation of cell surface-associated
plasminogen activation. J Biol Chem 1999;274:1527815283.
342. Piguet PF, Vesin C, Donati Y, et al. Urokinase receptor
(uPAR, CD87) is a platelet receptor important for kinetics and
TNF-induced endothelial adhesion in mice. Circulation
1999;99:33153321.
343. Jiang YP, Pannell R, Liu JN, et al. Evidence for a novel
binding protein to urokinase-type plasminogen activator in
platelet membranes. Blood 1996; 87:27752781.
P.361
344. Longstaff C, Merton RE, Fabregas P, et al. Characterization
of cell- associated plasminogen activation catalyzed by
urokinase-type plasminogen activator, but independent of
urokinase receptor (uPAR, CD87). Blood 1999;93:38393846.
345. Stump DC, Stassen JM, Demarsin E, et al. Comparative
thrombolytic properties of single-chain forms of urokinase-type
plasminogen activator. Blood 1987;69:592596.
346. Pannell R, Black J, Gurewich V. Complementary modes of
action of tissue-type plasminogen activator and pro-urokinase
by which their synergistic effect on clot lysis may be explained.
J Clin Invest 1988;81:853859.
347. Longstaff C, Clough AM, Gaffney PJ. Kinetics of plasmin
activation of single chain urinary-type plasminogen activator
(scuPA) and demonstration of a high affinity interaction
between scuPA and plasminogen. J Biol Chem
1992;267:173179.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

110 de 138

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

348. Lenich C, Pannell R, Gurewich V. The effect of the


carboxy-terminal lysine of urokinase on the catalysis of
plasminogen activation. Thromb Res 1991;64:6980.
349. Collen D, Stassen J-M, De Cock F. Synergistic effect of
thrombolysis of sequential infusion of tissue-type plasminogen
activator (tPA), single chain urokinase-type plasminogen
activator (scuPA) and urokinase in the rabbit jugular vein
thrombosis model. Thromb Haemost 1987;58:943946.
350. Hauert J, Nicoloso G, Schleuning WD, et al. Plasminogen
activators in dextran sulfate-activated euglobulin fractions: a
molecular analysis of factor XII- and prekallikrein-dependent
fibrinolysis. Blood 1989;73:994999.
351. Binnema DJ, Dooijewaard G, Iersel JJL, et al. The
contact-system dependent plasminogen activator from human
plasma: identification and characterization. Thromb Haemost
1990;64:390397.
352. Schousboe I, Feddersen K, Rjkjaer R. Factor XIIa is a
kinetically favorable plasminogen activator. Thromb Haemost
1999;82:10411046.
353. Besser D, Verde P, Nagamine Y, et al. Signal transduction
and the uPA/uPAR system. Fibrinolysis 1997;10:215237.
354. Irigoyen JP, Nagamine Y. Cytoskeletal reorganization leads
to induction of the urokinase-type plasminogen activator gene
by activating FAK and Src and subsequently the Ras/Erk
signaling pathway. Biochem Biophys Res Commun
1999;262:666670.
355. D'Orazio D, Besser D, Marksitzer R, et al. Cooperation of
two PEA3/AP1 sites in uPA gene induction by TPA and FGF-2.
Gene 1997;201:179187.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

111 de 138

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

356. De Cesare D, Palazzolo M, Berthelsen J, et al.


Characterization of UEF-4, a DNA-binding protein required for
transcriptional synergism between two AP-1 sites in the human
urokinase enhancer. J Biol Chem 1997;272: 2392123929.
357. Smicun Y, Kopf E, Miskin R. The 3-untranslated region of
the urokinase gene enhances the expression of chimeric genes
in cultured cells and correlates with specific brain expression in
transgenic mice. Eur J Biochem 1998;251:704715.
358. Nanbu R, Menoud PA, Nagamine Y. Multiple
instability-regulating sites in the 3 untranslated region of the
urokinase-type plasminogen activator mRNA. Mol Cell Biol
1994;14:49204928.
359. Singh I, Burnand KG, Collins M, et al. Failure of thrombus
to resolve in urokinase-type plasminogen activator
gene-knockout mice: rescue by normal bone marrow-derived
cells. Circulation 2003;107:869875.
360. Bugge TH, Flick MJ, Danton MJ, et al. Urokinase-type
plasminogen activator is effective in fibrin clearance in the
absence of its receptor or tissue-type plasminogen activator.
Proc Natl Acad Sci U S A 1996;93:58995904.
361. Carmeliet P, Moons L, Lijnen R, et al. Urokinase-generated
plasmin activates matrix metalloproteinases during aneurysm
formation. Nat Genet 1997;17:439444.
362. Carmeliet P, Moons L, Herbert J-M, et al. Urokinase but not
tissue type plasminogen activator mediates arterial neointima
formation in mice. Circ Res 1997;81:829839.
363. Carmeliet P, Moons L, Dewerchin M, et al.
Receptor-independent role of urokinase-type plasminogen
activator in pericellular plasmin and matrix metalloproteinase

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

112 de 138

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

proteolysis during vascular wound healing in mice. J Cell Biol


1998;140:233245.
364. Deindl E, Ziegelhoffer T, Kanse SM, et al.
Receptor-independent role of the urokinase-type plasminogen
activator during arteriogenesis. FASEB J 2003;17:11741176.
365. Gyetko MR, Aizenberg D, Mayo-Bond L. Urokinase-deficient
and urokinase receptor-deficient mice have impaired neutrophil
antimicrobial activation in vitro. J Leukoc Biol
2004;76:648356.
366. Piguet PF, Vesin C, Donati Y, et al. Urokinase receptor
(uPAR, CD87) is a platelet receptor important for kinetics and
TNF-induced endothelial adhesion in mice. Circulation
1999;99:33153321.
367. Gu JM, Johns A, Morser J, et al. Urokinase plasminogen
activator receptor promotes macrophage infiltration into the
vascular wall of ApoE deficient mice. J Cell Physiol 2005;
204:7382.
368. Jo M, Thomas KS, Wu L, et al. Soluble urokinase-type
plasminogen activator receptor inhibits cancer cell growth and
invasion by direct urokinase-independent effects on cell
signaling. J Biol Chem 2003;278:4669246698.
369. Tillett WS, Garner RL. The fibrinolytic activity of hemolytic
streptococci. J Exp Med 1933;68:485502.
370. Gladysheva IP, Turner RB, Sazonova IY, et al.
Coevolutionary patterns in plasminogen activation. Proc Natl
Acad Sci U S A 2003;100:91689172.
371. Wohl RC, Summaria L, Arzadon L, et al. Steady state
kinetics of activation of human and bovine plasminogens by

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

113 de 138

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

streptokinase and its equimolar complexes with various


activated forms of human plasminogen. J Biol Chem
1978;253:14021407.
372. Parry MA, Zhang XC, Bode I. Molecular mechanisms of
plasminogen activation: bacterial cofactors provide clues.
Trends Biochem Sci 2000;25: 5359.
373. Cederholm-Williams SA, De Cock F, Lijnen HR, et al.
Kinetics of the reaction between streptokinase, plasmin and
2 -antiplasmin. Eur J Biochem 1979;100:125132.
374. Mundada LV, Prorok M, DeFord ME, et al.
Structure-function analysis of the streptokinase amino terminus
(residues 159). J Biol Chem 2003;278: 2442124427.
375. Conejerolara F, Parrado J, Azuaga AI, et al. Thermal
stability of the three domains of streptokinase studied by
circular dichroism and nuclear magnetic resonance. Protein Sci
1996;5:25832591.
376. Wang X, Lin X, Loy JA, et al. Crystal structure of the
catalytic domain of human plasmin complexed with
streptokinase. Science 1998;281: 16621665.
377. Wang X, Tang J, Hunter B, et al. Crystal structure of
streptokinase -domain. FEBS Lett 1999;459:8589.
378. Chaudhary A, Vasudha S, Rajagopal K, et al. Function of
the central domain of streptokinase in substrate plasminogen
docking and processing revealed by site-directed mutagenesis.
Protein Sci 1999;8:27912805.
379. Reddy KNN. Streptokinase: biochemistry and clinical
application. Enzyme 1988;40:7989.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

114 de 138

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

380. Longstaff C, Whitton CM. A proposed reference method for


plasminogen activators that enables calculation of enzyme
activities in SI units. J Thromb Haemost 2004;2:14161421.
381. Bachmann F. Development of antibodies against perorally
and rectally administered streptokinase in man. J Lab Clin Med
1968;72:228238.
382. Ojalvo AG, Pozo L, Labarta V, et al. Prevalence of
circulating antibodies against a streptokinase C-terminal
peptide in normal blood donors. Biochem Biophys Res Commun
1999;263:454459.
383. Gonias SL, Einarsson M, Pizzo SV. Catabolic pathways for
streptokinase, plasmin and streptokinase activator complex in
mice: in vivo reaction of plasminogen activator with
2 macroglobulin. J Clin Invest 1982; 7:412423.
384. Hibbs MJ, Fears R, Ferres H, et al. Determination of the
deacetylation rate of p-anisoyl plasminogen streptokinase
activator complex (APSAC, EMINASE) in human plasma, blood
and plasma clots. Fibrinolysis 1989;3:235240.
385. Fears R. Development of anisoylated
plasminogen-streptokinase activator complex from the acyl
enzyme concept. Semin Thromb Hemost 1989;15: 129139.
386. Lack CH. Staphylokinase: an activator of plasma protease.
Nature 1948; 161:559560.
387. Lewis JH, Ferguson JH. A proteolytic enzyme sytem of the
blood. III. Activation of dog serum profibrinolysin by
staphylokinase. Am J Physiol 1951;166:594603.
388. Collen D. Staphylokinase: a potent, uniquely

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

115 de 138

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

fibrin-selective thrombolytic agent. Nat Med 1998;4:279284.


389. Laroche Y, Heymans S, Capaert S, et al. Recombinant
staphylokinase variants with reduced antigenicity due to
elimination of B-lymphocyte epitopes. Blood
2000;96:14251432.
390. Vanwetswinkel S, Plaisance S, Zhi-Yong Z, et al.
Pharmacokinetic and thrombolytic properties of cysteine-linked
polyethylene glycol derivatives of staphylokinase. Blood
2000;95:936942.
391. Vanderschueren S, Barrios L, Kerdsinchai P et al, The
STAR Trial Group. A randomized trial of recombinant
staphylokinase versus alteplase for coronary artery patency in
acute myocardial infarction. Circulation 1995; 92:20442049.
392. Heymans S, Vanderschueren S, Verhaeghe R, et al.
Outcome and one year follow-up of intra-arterial staphylokinase
in 191 patients with peripheral arterial occlusion. Thromb
Haemost 2000;83:666671.
393. Hawkey C. Plasminogen activator in the saliva of the
vampire bat Desmodus rotundus. Nature 1966;211:434435.
394. Gardell SJ, Duong LT, Diehl RE, et al. Isolation,
characterization and cDNA cloning of a vampire bat salivary
plasminogen activator. J Biol Chem 1989;264:1794717952.
395. Krtzschmar J, Haendler B, Langer G, et al. The
plasminogen activator family from the salivary gland of the
vampire bat Desmodus rotundus: cloning and expression. Gene
1991;105:229237.
396. Schleuning W-D, Donner P. Desmodus rotundus (common
vampire bat) salivary plasminogen activator. In: Bachmann F,

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

116 de 138

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

ed. Fibrinolytics and antifibrinolytics, handbook of experimental


pharmacology. Heidelberg: Springer, 2000:447468.
397. Bergum PW, Gardell SJ. Vampire bat salivary plasminogen
activator exhibits a strict and fastidious requirement for
polymeric fibrin as its cofactor, unlike human tissue-type
plasminogen activator. A kinetic analysis. J Biol Chem
1992;267:1772617731.
398. Bringmann P, Gruber D, Liese A, et al. Structural features
mediating fibrin selectivity of vampire bat plasminogen
activators. J Biol Chem 1995;270: 2559625603.
399. Stewart RJ, Fredenburgh JC, Weitz JI. Characterization of
the interactions of plasminogen and tissue and vampire bat
plasminogen activators with fibrinogen, fibrin, and the complex
of D-dimer noncovalently linked to fragment E. J Biol Chem
1998;273:1829218299.
400. Renatus M, Stubbs MT, Huber R, et al. Catalytic domain
structure of vampire bat plasminogen activator: a molecular
paradigm for proteolysis without activation cleavage.
Biochemistry 1997;36:1348313493.
P.362
401. Montoney M, Gardell SJ, Marder VJ. Comparison of the
bleeding potential of vampire bat salivary plasminogen activator
versus tissue plasminogen activator in an experimental rabbit
model. Circulation 1995;91: 15401544.
402. Witt W, Maass B, Baldus B, et al. Coronary thrombolysis
with Desmodus salivary plasminogen activator in dogs. Fast and
persistent recanalization by intravenous bolus administration.
Circulation 1994;90:421426.
403. Hildebrand M, Bhargava AS, Bringmann P, et al.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

117 de 138

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

Pharmacokinetics of the novel plasminogen activator Desmodus


rotundus plasminogen activator in animals and extrapolation to
man. Fibrinolysis 1996;10:269276.
404. Liberatore GT, Samson A, Bladin C, et al. Vampire bat
salivary plasminogen activator (desmoteplase): a unique
fibrinolytic enzyme that does not promote neurodegeneration.
Stroke 2003;34:537543.
405. Collen D, Wiman B. Fast-acting plasmin inhibitor in human
plasma. Blood 1978;51:563569.
406. Mast AE, Enghild JJ, Pizzo SV, et al. Analysis of the plasma
elimination kinetics and conformational stabilities of native,
proteinase-complexed, and reactive site cleaved serpins:
comparison of 2 -proteinase inhibitor, 2 -antichymotrypsin,
antithrombin III, 2 -antiplasmin, angiotensinogen, and
ovalbumin. Biochemistry 1991;30:17231730.
407. Holmes WE, Nelles L, Lijnen HR, et al. Primary structure of
human 2 -antiplasmin a serine protease inhibitor (serpin). J
Biol Chem 1987;262: 16591664.
408. Hirosawa S, Nakamura Y, Miura O, et al. Organization of
the human 2 -plasmin inhibitor gene. Proc Natl Acad Sci U S A
1988;85:68366840.
409. Kato A, Hirosawa S, Toyota S, et al. Localization of the
human 2 -plasmin inhibitor gene (PLI) to 17p13. Cytogenet Cell
Genet 1993;62: 190191.
410. Lee KN, Tae W-C, Jackson KW, et al. Characterization of
wild-type and mutant 2 -antiplasmin fibrinolysis enhancement
by reactive site mutant. Blood 1999;94:164171.
411. Wiman B, Boman L, Collen D. On the kinetics of the

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

118 de 138

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

reaction between human antiplasmin and a


low-molecular-weight form of plasmin. Eur J Biochem
1978;87:143146.
412. Sumi Y, Ichikawa Y, Nakamura Y, et al. Expression and
characterization of pro 2 -plasmin inhibitor. J Biochem
1989;106:703707.
413. Bangert K, Johnsen AH, Christensen U et al. Different
N-terminal forms of 2 -plasmin inhibitor in human plasma.
Biochem J 1993;291:523625.
414. Koyama T, Koike Y, Toyota S, et al. Different 2 -plasmin
form with 12 additional residues of 2 -plasmin inhibitor from
human plasma and culture media of Hep G2 cells. Biochem
Biophys Res Commun 1994;200: 417422.
415. Lee KN, Jackson KW, Christiansen VJ, et al. A novel
plasma proteinase potentiates 2 -antiplasmin inhibition of fibrin
digestion. Blood 2004;103: 37833788.
416. Kimura S, Aoki N. Cross-linking site in fibrinogen for
2 -plasmin inhibitor. J Biol Chem 1986;261:1559115595.
417. Booth NA. Regulation of fibrinolytic activity by localization
of inhibitors to fibrin(ogen). Fibrinolysis Proteolysis
2000;14:206213.
418. Lee KN, Tae W-C, Jackson KW, et al. Characterization of
wild-type and mutant 2 -antiplasmin: fibrinolysis enhancement
by reactive site mutant. Blood 1999;94:164171.
419. Reed GL, Matsueda GR, Haber E. Synergistic fibrinolysis:
combined effects of plasminogen activators and an antibody
that inhibits 2 -antiplasmin. Proc Natl Acad Sci U S A
1990;87:11141118.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

119 de 138

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

420. Clemmensen I, Thorsen S, Mllertz S, et al. Properties of


three different molecular forms of 2 -plasmin inhibitor. Eur J
Biochem 1981;120: 105122.
421. Sasaki T, Morita T, Iwanaga S. Identification of the
plasminogen-binding site of human 2 -plasmin inhibitor. J
Biochem 1986;99:16961705.
422. Kluft C, Los N. Demonstration of two forms of
2 -antiplasmin in plasma by modified crossed
immunoelectrophoresis. Thromb Res 1981;21: 6571.
423. Collen D. On the regulation and control of fibrinolysis.
Thromb Haemost 1980;43:7789.
424. Wiman B, Collen D. On the kinetics of the reaction
between human antiplasmin and plasmin. Eur J Biochem
1978;84:573578.
425. Frank PS, Douglas JT, Locher M, et al.
Structural/functional characterization of the 2 -plasmin
inhibitor C-terminal peptide. Biochemistry 2003;42: 10781085.
426. Wang H, Yu A, Wiman B, et al. Identification of amino
acids in antiplasmin involved in its noncovalent lysine-binding
site-dependent interaction with plasmin. Eur J Biochem
2003;270:20232029.
427. Paramo JA, Gascoine PS, Pring JB, et al. The relative
inhibition by 2 -antiplasmin and plasminogen activator
inhibitor-1 of clot lysis in vitro. Fibrinolysis 1990;4:169175.
428. Robbie LA, Booth NA, Croll AM, et al. The roles of
2 -antiplasmin and plasminogen activator inhibitor 1 (PAI-1) in
the inhibition of clot lysis. Thromb Haemost 1993;70:301306.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

120 de 138

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

429. Robbie LA, Bennett B, Keyt BA, et al. Effective lysis of


model thrombi by a t-PA mutant (A473S) that is resistant to
2 -antiplasmin. Br J Haematol 2000;111:517523.
430. Koie K, Kamiya T, Ogata K, et al. 2 -plasmin-inhibitor
deficiency (Miyasato disease). Lancet 1978;2:13341336.
431. Aoki N, Saito H, Kamiya T, et al. Congenital deficiency of
2 -plasmin inhibitor associated with severe hemorrhagic
tendency. J Clin Invest 1979; 63:877884.
432. Rijken DC, Groeneveld E, Kluft C, et al. 2 -Antiplasmin
Enschede is not an inhibitor, but a substrate, of plasmin.
Biochem J 1988;255:609615.
433. Holmes WE, Kijnen HR, Nelles L, et al. 2 -Antiplasmin
Enschede: alanine insertion and abolition of plasmin inhibitory
activity. Science 1987;238: 209211.
434. Lijnen HR, Okada K, Matsuo O, et al. 2 -Antiplasmin gene
deficiency in mice is associated with enhanced fibrinolytic
potential without overt bleeding. Blood 1999;93:22742281.
435. Okada K, Ueshima S, Imano M, et al. The regulation of
liver regeneration by the plasmin/ 2 -antiplasmin system. J
Hepatol 2004;40:110116.
436. Takei M, Matsuno H, Okada K, et al. Lack of 2 -antiplasmin
enhances ADP induced platelet micro-aggregation through the
presence of excess active plasmin in mice. J Thromb
Thrombolysis 2002;14:205211.
437. Matsuno H, Kozawa O, Yoshimi N, et al. Lack of
2 -antiplasmin promotes pulmonary heart failure via
overrelease of VEGF after acute myocardial infarction. Blood

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

121 de 138

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

2002;100:24872493.
438. Sakurai Y, Takahashi T, Arakawa H, et al. Trypsin-like
endopeptidase(s) naturally entrapped in human blood
2 -macroglobulin. Biomed Res 1996; 17:347350.
439. Sottrup-Jensen L, Stepanik TM, Kristensen T, et al.
Primary structure of human 2 -macroglobulin. V. The complete
structure. J Biol Chem 1984; 259:83188327.
440. Kan CC, Solomon E, Belt KT, et al. Nucleotide sequence of
cDNA encoding human 2 -macroglobulin and assignment of the
chromosomal locus. Proc Natl Acad Sci U S A
1985;82:22822286.
441. Travis J, Salvesen GS. Human plasma proteinase
inhibitors. Annu Rev Biochem 1983;52:655709.
442. Kolodziej SJ, Klueppelberg HU, Nolasco N, et al.
Three-dimensional structure of the human plasmin
2 -macroglobulin. J Struct Biol 1998;123:124133.
443. 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.
444. 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.
445. Declerck PJ, De Mol M, Alessi M-C, et al. Purification and
characterization of a plasminogen activator inhibitor 1 binding
protein from human plasma. Identification as a multimeric form
of S protein (vitronectin). J Biol Chem 1988;263:1545415461.
446. Kruithof EKO, Gudinchet A, Bachmann F. Plasminogen

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

122 de 138

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

activator inhibitor 1 and plasminogen activator inhibitor 2 in


various disease states. Thromb Haemost 1988;59:712.
447. Booth NA, Simpson AJ, Croll A, et al. Plasminogen
activator inhibitor (PAI-1) in plasma and platelets. Br J
Haematol 1988;70:327333.
448. Declerck PJ, Alessi M-C, Verstreken MV, et al.
Measurement of plasminogen activator inhibitor 1 in biologic
fluids with a murine monoclonal antibody-based enzyme-linked
immunosorbent assay. Blood 1988;71:220225.
449. Johnson SA, Schneider CL. The existence of antifibrinolysin
activity in platelets. Science 1953;117:229230.
450. Konkle BA, Schick PK, He X, et al. Plasminogen activator
inhibitor-1 mRNA is expressed in platelets and megakaryocytes
and the megakaryoblastic cell line CHRF-288. Arterioscler
Thromb 1993;13:669674.
451. van Mourik JA, Lawrence DA, Loskutoff DJ. Purification of
an inhibitor of plasminogen activator (antiactivator) synthesized
by endothelial cells. J Biol Chem 1984;259:1491414921.
452. Cwikel BJ, Barouski-Miller PA, Coleman PL, et al.
Dexamethasone induction of an inhibitor of plasminogen
activator in HTC hepatoma cells. J Biol Chem
1984;259:68476851.
453. Morange PE, Alessi MC, Verdier M, et al. PAI-1 produced ex
vivo by human adipose tissue is relevant to PAI-1 blood level.
Arterioscler Thromb Vasc Biol 1999;19:13611365.
454. Loskutoff DJ, Samad F. The adipocyte and hemostatic
balance in obesity: studies of PAI-1. Arterioscler Thromb Vasc
Biol 1998;18:16.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

123 de 138

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

455. Crandall DL, Quinet EM, Morgan GA, et al. Synthesis and
secretion of plasminogen activator inhibitor-1 by human
preadipocytes. J Clin Endocrinol Metab 1999;84:32223227.
456. Loskutoff DJ. Regulation of PAI-1 gene expression.
Fibrinolysis 1991;5: 197206.
457. Booth NA. The natural inhibitors of fibrinolysis. In: Bloom
AL, Forbes CD, Thomas DP, et al., eds. Haemostasis and
thrombosis, 3rd ed. Edinburgh: Churchill Livingstone,
1994:699717.
458. Robbie LA, Bennett B, Croll AM, et al. Proteins of the
fibrinolytic system in human thrombi. Thromb Haemost
1996;75:127133.
459. Potter van Loon BJ, Rijken DC, Brommer EJP, et al. The
amount of plasminogen, tissue-type plasminogen activator and
plasminogen activator inhibitor type 1 in human thrombin and
relation to ex vivo lysability. Thromb Haemost
1992;67:101105.
460. Kruithof EKO.The inhibitors of the fibrinolytic system. In:
Bachmann F, ed. Fibrinolytics and antifibrinolytics. Handbook of
experimental pharmacology. Vol 146. Berlin, Springer,
2001:113139.
461. Dival J, Nguyen G, Gross S, et al. A lifelong bleeding
disorder associated with a deficiency of plasminogen activator
inhibitor type 1. Blood 1991; 77:528532.
462. 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.
P.363

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

124 de 138

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

463. Lee MH, Vosburgh E, Anderson K, et al. Deficiency of


plasma plasminogen activator inhibitor 1 results in
hyperfibrinolytic bleeding. Blood 1993; 81:23572362.
464. 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.
465. Carmeliet P, Stassen JM, Schoonjans L, et al. Plasminogen
activator inhibitor-1 gene-deficient mice. II. Effects on
hemostasis, thrombosis and thrombolysis. J Clin Invest
1993;92:17562760.
466. 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.
467. Eitzman DT, Westrick RJ, Nabel EG, et al. Plasminogen
activator inhibitor-1 and vitronectin promote vascular
thrombosis in mice. Blood 2000;95:577580.
468. Farrehi PM, Ozaki CK, Carmeliet P, et al. Regulation of
arterial thrombolysis by plasminogen activator inhibitor-1 in
mice. Circulation 1998;97: 10021008.
469. 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.
470. Bachmann F. The role of plasminogen activator inhibitor
type 1 (PAI-1) in the clinical setting, including deep vein
thrombosis. In: Glas-Greenwalt P, ed. Fibrinolysis in disease.
Molecular and hemovascular aspects of fibrinolysis. Boca Raton,

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

125 de 138

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

FL: CRC Press, 1995:7986.


471. Vaughan DE. Plasminogen activator inhibitor-1: a common
denominator in cardiovascular disease. J Investig Med
1998;46:370376.
472. Juhan-Vague I, Pyke SDM, Alessi MC, et al. Fibrinolytic
factors and the risk of myocardial infarction or sudden death in
patients with angina pectoris. Circulation 1996;94:20572063.
473. Carroll VA, Binder BR. The role of the plasminogen
activation system in cancer. Semin Thromb Hemost
1999;25:183197.
474. Scarabin P-Y, Aillaud M-F, Amouyel P et al, The PRIME
Study. Associations of fibrinogen, factor VII and PAI-1 with
baseline findings among 10,500 male participants in a
prospective study of myocardial infarction. Thromb Haemost
1998;80:749756.
475. Dawson S, 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.
476. Eriksson P, Kallin B, Van'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.
477. Wiman B, Hamsten A. Impaired fibrinolysis and risk of
thromboembolism. Prog Cardiovasc Dis 1991;34:179192.
478. Lowe GDO, Yarnell JWG, Sweetnam PM, et al. Fibrin
D-dimer, tissue plasminogen activator, plasminogen activator
inhibitor, and the risk of major ischaemic heart disease in the

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

126 de 138

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

Caerphilly study. Thromb Haemost 1998;79:129133.


479. Pralong G, Calandra T, Glauser MP, et al. Plasminogen
activator inhibitor 1: a new prognostic marker in septic shock.
Thromb Haemost 1989;61: 459462.
480. Arts J, Kockx M, Princen HMG, et al. Studies on the
mechanism of fibrate-inhibited expression of plasminogen
activator inhibitor-1 in cultured hepatocytes from cynomolgus
monkey. Arterioscler Thromb Vasc Biol 1997;17:2632.
481. Murakami J, Ohtani A, Murata S. Protective effect of
T-686, an inhibitor of plasminogen activator inhibitor-1
production, against the lethal effect of lipopolysaccharide in
mice. Jpn J Pharmacol 1997;75:291294.
482. Ohtani A, Murakami J, Hirano-Wakimoto A. T-686, a novel
inhibitor of plasminogen activator inhibitor-1, inhibits
thrombosis without impairment of hemostasis in rats. Eur J
Pharmacol 1997;330:151156.
483. Vinogradsky B, Bell SP, Woodcock-Mitchell J, et al. A new
butadiene derivative, T-686, inhibits plasminogen activator
inhibitor type-1 production in vitro by cultured human vascular
endothelial cells and development of atherosclerotic lesions in
vivo in rabbits. Thromb Res 1997;85:305314.
484. Eitzman DT, Fay WP, Lawrence DA, et al. Peptide-mediated
inactivation of recombinant and platelet plasminogen activator
inhibitor-1 in vitro. J Clin Invest 1995;95:24162420.
485. Neve J, Leone PA, Carroll AR, et al. Sideroxylonal C, a new
inhibitor of human plasminogen activator inhibitor type-1, from
the flowers of Eucalyptus albens. J Nat Prod 1999;62:324326.
486. Bjrquist P, Ehnebom J, Inghardt T, et al. Identification of

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

127 de 138

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

the binding site for a low-molecular-weight inhibitor of


plasminogen activator inhibitor type 1 by site-directed
mutagenesis. Biochemistry 1998;37:12271234.
487. Charlton PA, Faint RW, Bent F, et al. Evaluation of a low
molecular weight modulator of human plasminogen activator
inhibitor-1 activity. Thromb Haemost 1996;75:808815.
488. Friederich PW, Levi M, Biemond BJ, et al. Novel
low-molecular-weight inhibitor of PAI-1 (XR5118) promotes
endogenous fibrinolysis and reduces postthrombolysis thrombus
growth in rabbits. Circulation 1997;96: 916921.
489. Einholm AP, Pedersen KE, Wind T, et al. Biochemical
mechanism of action of a diketopiperazine inactivator of
plasminogen activator inhibitor-1. Biochem J
2003;373:723732.
490. Ye RD, Ahern SM, Le Beau MM, et al. Structure of the gene
for human plasminogen activator inhibitor-2. The nearest
mammalian homologue of chicken ovalbumin. J Biol Chem
1989;264:54955502.
491. Bachmann F. The enigma PAI-2. Gene expression,
evolutionary and functional aspects. Thromb Haemost
1995;74:172179.
492. Kruithof EK, Baker MS, Bunn CL. Biological and clinical
aspects of plasminogen activator inhibitor type 2. Blood
1995;86:40074024.
493. Kawano T, Morimoto K, Uemura Y. Partial purification and
properties of urokinase inhibitor from human placenta. J
Biochem (Tokyo) 1970;67: 333342.
494. stedt B, Lecander I, Brodin T, et al. Purification of a

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

128 de 138

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

specific placental plasminogen activator inhibitor by monoclonal


antibody and its complex formation with plasminogen
activation. Thromb Haemost 1985;53:122125.
495. Kruithof EKO, Vassalli J-D, Schleuning W-D, et al.
Purification and characterization of a plasminogen activator
inhibitor from the histiocytic lymphoma cell line U-937. J Biol
Chem 1986;261:1120711213.
496. Risse BC, Brown H, Lavker RM, et al. Differentiating cells
of murine stratified squamous epithelia constitutively express
plasminogen activator inhibitor type 2 (PAI-2). Histochem Cell
Biol 1998;110:559569.
497. Jensen PH, Lorand L, Ebbesen P, et al. Type-2
plasminogen-activator inhibitor is a substrate for trophoblast
transglutaminase and factor XIII 2 . Transglutaminase-catalyzed
cross-linking to cellular and extracellular structures. Eur J
Biochem 1993;214:141146.
498. Schwartz BS, Bradshaw JD. Differential regulation of tissue
factor and plasminogen activator inhibitor by human
mononuclear cells. Blood 1989; 74:16441650.
499. Ritchie H, Robbie LA, Kinghorn S, et al. Monocyte
plasminogen activator inhibitor 2 (PAI-2) inhibits
u-PA-mediated fibrinolysis and is cross-linked to fibrin. Thromb
Haemost 1999;81:96103.
500. Genton C, Kruithof EK, Schleuning WD. Phorbol ester
induces the biosynthesis of glycosylated and nonglycosylated
plasminogen activator inhibitor 2 in high excess over
urokinase-type plasminogen activator in human U-937
lymphoma cells. J Cell Biol 1987;104:705712.
501. Ritchie H, Booth NA. Secretion of plasminogen activator

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

129 de 138

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

inhibitor 2 by human peripheral blood monocytes occurs via an


endoplasmic reticulum-golgi-independent pathway. Exp Cell Res
1998;242:439450.
502. Harrop SJ, Jankova L, Coles M, et al. The crystal structure
of plasminogen activator inhibitor 2 at 2.0 resolution:
implications for serpin function. Structure Fold Des
1999;7:4354.
503. Jensen PH, Schler E, Woodrow G, et al. A unique
interhelical insertion in plasminogen activator inhibitor-2
contains three glutamines, Gln 83 , Gln 84 , Gln 86 , essential for
transglutaminase-mediated cross-linking. J Biol Chem
1994;269:1539415398.
504. Ritchie H, Lawrie LC, Crombie PW, et al. Cross-linking of
plasminogen activator inhibitor 2 and 2 -antiplasmin to
fibrin(ogen). J Biol Chem 2000; 275:2491524920.
505. Jensen PH, Jensen TG, Laug WE, et al. The exon 3 encoded
sequence of the intracellular serine proteinase inhibitor
plasminogen activator inhibitor 2 is a protein binding domain. J
Biol Chem 1996;271:2689226899.
506. Mikus P, Urano T, Liljestrm P, et al.
Plasminogen-activator inhibitor type 2 (PAI-2) is a
spontaneously polymerising SERPINbiochemical
characterisation of the recombinant intracellular and
extracellular forms. Eur J Biochem 1993;218:10711082.
507. Wilczynska M, Lobov S, Ohlsson PI, et al. A
redox-sensitive loop regulates plasminogen activator inhibitor
type 2 (PAI-2) polymerization. EMBO J 2003;22:17531761.
508. Lobov S, Wilczynska M, Bergstrom F, et al. Structural
bases of the redox-dependent conformational switch in the

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

130 de 138

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

serpin PAI-2. J Mol Biol 2004; 344:13591368.


509. Kruithof EKO, Tran-Thang C, Gudinchet A, et al.
Fibrinolysis in pregnancy: a study of plasminogen activator
inhibitors. Blood 1987;69:460466.
510. Bonnar J, Daly L, Sheppard BL. Changes in the fibrinolytic
system during pregnancy. Sem Thromb Hemostas
1990;16:221229.
511. Reith A, Booth NA, Moore NR, et al. Plasminogen activator
inhibitors (PAI-1 and PAI-2) in normal pregnancies,
pre-eclampsia and hydatididform mole. Br J Obstet Gynaecol
1993;100:370374.
512. Grancha S, Estells A, Gilabert J, et al. Decreased
expression of PAI-2 mRNA and protein in pregnancies
complicated with intrauterine fetal growth retardation. Thromb
Haemost 1996;76:761767.
513. Booth NA, Reith A, Bennett B. A plasminogen activator
inhibitor (PAI-2) circulates in two molecular forms during
pregnancy. Thromb Haemost 1988;59:7779.
514. Scherrer A, Kruithof EKO, Grob J-P. Plasminogen activator
inhibitor-2 in patients with monocytic leukemia. Leukemia
1991;5:479486.
515. Robbie LA, Dummer S, Booth NA, et al. Plasminogen
activator inhibitor 2 and urokinase-type plasminogen activator
in plasma and leucocytes in patients with severe sepsis. Br J
Haematol 2000;109:342348.
516. Antalis TM, La Linn M, Donnan K, et al. The serine
proteinase inhibitor (serpin) plasminogen activation inhibitor
type 2 protects against viral cytopathic effects by constitutive

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

131 de 138

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

interferon / priming. J Exp Med 1998; 187:17991811.


517. Darnell GA, Antalis TM, Johnstone RW, et al. Inhibition of
retinoblastoma protein degradation by interaction with the
serpin plasminogen activator inhibitor 2 via a novel consensus
motif. Mol Cell Biol 2003;23:65206532.
518. Mueller BM, Yu YB, Laug WE. Overexpression of
plasminogen activator inhibitor 2 in human melanoma cells
inhibits spontaneous metastasis in scid/scid mice. Proc Natl
Acad Sci U S A 1995;92:205209.
P.364
519. Varro A, Noble PJ, Pritchard DM, et al. Helicobacter pylori
induces plasminogen activator inhibitor 2 in gastric epithelial
cells through nuclear factor-kappaB and RhoA: implications for
invasion and apoptosis. Cancer Res 2004;64:16951702.
520. Dougherty KM, Pearson JM, Yang AY, et al. The
plasminogen activator inhibitor-2 gene is not required for
normal murine development or survival. Proc Natl Acad Sci U S
A 1999;96:686691.
521. Salvesen GS, Catanese JJ, Kress LF, et al. Primary
structure of the reactive site of human C1-inhibitor. J Biol Chem
1985;260:24322436.
522. Tosi M, Duponchel C, Bourgarel P, et al. Molecular cloning
of human C1 inhibitor: sequence homologies with 1 -antitrypsin
and other members of the serpins superfamily. Gene
1986;42:265272.
523. Bock SC, Skriver K, Nielsen E, et al. Human C1 inhibitor:
primary structure, cDNA cloning, and chromosomal localization.
Biochemistry 1986; 25:42924301.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

132 de 138

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

524. Huisman LGM, Van Griensven JMT, Kluft C. On the role of


C1-inhibitor as inhibitor of tissue-type plasminogen activator in
human plasma. Thromb Haemost 1995;73:466471.
525. Koide T, Foster D, Yoshitake S, et al. Amino acid sequence
of human histidine-rich glycoprotein derived from the nucleotide
sequence of its cDNA. Biochemistry 1986;25:22202225.
526. Lijnen HR, Hoylaerts M, Collen D. Isolation and
characterization of a human plasma protein with affinity for the
lysine binding sites in plasminogen. J Biol Chem
1980;255:1021410222.
527. Ichinose A, Mimuro J, Koide T, et al. Histidine-rich
glycoprotein and 2 -plasmin in inhibition of plasminogen
binding to fibrin. Thromb Res 1984;33:401407.
528. Jones AL, Hulett MD, Altin JG, et al. Plasminogen is
tethered with high affinity to the cells surface by the plasma
protein, histidine-rich glycoprotein. J Biol Chem
2004;279:3826738276.
529. Shigekiyo T, Ohshima T, Oka H, et al. Congenital
histidine-rich glycoprotein deficiency. Thromb Haemost
1993;70:263265.
530. Nesheim M. Fibrinolysis and the plasma carboxypeptidase.
Curr Opin Hematol 1998;5:309313.
531. Bajzar L, Nesheim M, Morser J, et al. Both cellular and
soluble forms of thrombomodulin inhibit fibrinolysis by
potentiating the activation of thrombin-activable fibrinolysis
inhibitor. J Biol Chem 1998;273:27922798.
532. Bouma BN, Marx PF, Mosnier LO, et al.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

133 de 138

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

Thrombin-activatable fibrinolysis inhibitor (TAFI, plasma


procarboxypeptidase B, procarboxypeptidase R,
procarboxypeptidase U). Thromb Res 2001;101:329354.
533. Bajzar L, Morser J, Nesheim M. TAFI, or plasma
procarboxypeptidase B, couples the coagulation and fibrinolytic
cascades through the thrombin-thrombomodulin complex. J Biol
Chem 1996;271:1660316608.
534. Leurs J, Wissing BM, Nerme V, et al. Different mechanisms
contribute to the biphasic pattern of carboxypeptidase U
(TAFIa) generation during in vitro clot lysis in human plasma.
Thromb Haemost 2003;89:264271.
535. Mutch NJ, Moore NR, Wang E, et al. Thrombus lysis by
uPA, scuPA and tPA is regulated by plasma TAFI. J Thromb
Haemost 2003;1:20002007.
536. Redlitz A, Nicolini FA, Malycky JL, et al. Inducible
carboxypeptidase activity. A role in clot lysis in vivo. Circulation
1996;93:13281330.
537. Nagashima M, Werner M, Wang M, et al. An inhibitor of
activated thrombin-activatable fibrinolysis inhibitor potentiates
tissue-type plasminogen activator-induced thrombolysis in a
rabbit jugular vein thrombolysis model. Thromb Res
2000;98:333342.
538. Klement P, Liao P, Bajzar L. A novel approach to arterial
thrombolysis. Blood 1999;94:27352743.
539. Nagashima M, Yin ZF, Zhao L, et al. Thrombin-activatable
fibrinolysis inhibitor (TAFI) deficiency is compatible with murine
life. J Clin Invest 2002;109:101110.
540. Silveira A, Schatteman K, Goossens F, et al. Plasma

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

134 de 138

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

procarboxypeptidase U in men with symptomatic coronary


artery disease. Thromb Haemost 2000;84:364368.
541. Morange PE, Aillaud MF, Nicaud V, et al. Ala147Thr and C
+ 1542G polymorphisms in the TAFI gene are not associated
with a higher risk of venous thrombosis in FV Leiden carriers.
Thromb Haemost 2001;86: 15831584.
542. Gils A, Alessi MC, Brouwers E, et al. Development of a
genotype 325-specific proCPU/TAFI ELISA. Arterioscler Thromb
Vasc Biol 2003;23: 11221127.
543. Guimaraes AH, van Tilburg NH, Vos HL, et al. Association
between thrombin activatable fibrinolysis inhibitor genotype and
levels in plasma: comparison of different assays. Br J Haematol
2004;124:659665.
544. Bouma BN, Meijers JCM. Fibrinolysis and the contact
system: a role for factor XI in the down-regulation of
fibrinolysis. Thromb Haemost 1999; 82:243250.
545. Broze GJ, Higuchi DA. Coagulation-dependent inhibition of
fibrinolysis role of carboxypeptidase-U and the premature lysis
of clots from hemophilic plasma. Blood 1996;88:38153823.
546. Mosnier LO, Lisman T, van den Berg HM, et al. The
defective down regulation of fibrinolysis in haemophilia A can
be restored by increasing the TAFI plasma concentration.
Thromb Haemost 2001;86:10351039.
547. Minnema MC, Friederich PW, Levi M, et al. Enhancement of
rabbit jugular vein thrombolysis by neutralization of factor XI.
In vivo evidence for a role of factor XI as an anti-fibrinolytic
factor. J Clin Invest 1998;101: 1014.
548. Scanu AM, Edelstein C. Kringle-dependent structural and

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

135 de 138

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

functional polymorphism of apolipoprotein(a). Biochim Biophys


Acta 1995;1256:112.
549. Lindahl G, Gersdorf E, Menzel HJ, et al. The gene for the
Lp(a)-specific glycoprotein is closely linked to the gene for
plasminogen on chromosome 6. Hum Genet 1989;81:149152.
550. Byrne CD, Schwartz K, Meer K, et al. The human
apolipoprotein(a)/ plasminogen gene cluster contains a novel
homologue transcribed in liver. Arterioscler Thromb
1994;14:534541.
551. Miles LA, Fles GM, Levin EG, et al. A potential basis for the
thrombotic risks associated with lipoprotein(a). Nature
1989;339:301303.
552. Hajjar KA, Gavish D, Breslow JL, et al. Lipoprotein(a)
modulation of endothelial cell surface fibrinolysis and its
potential role in atherosclerosis. Nature 1989;339:303305.
553. Gonzalez-Gronow M, Edelberg JM, Pizzo SV. Further
characterization of the cellular plasminogen binding site:
evidence that plasminogen 2 and lipoprotein(a) compete for the
same site. Biochemistry 1989;28:23742377.
554. Ezratty A, Simon DI, Loscalzo J. Lipoprotein(a) binds to
human platelets and attenuates plasminogen binding and
activation. Biochemistry 1993; 32:46284633.
555. Plow EF, Herren T, Redlitz A, et al. The cell biology of the
plasminogen system. FASEB J 1995;9:939945.
556. Sangrar W, Gabel BR, Boffa MB, et al. The solution phase
interaction between apolipoprotein(a) and plasminogen inhibits
the binding of plasminogen to a plasmin-modified fibrinogen
surface. Biochemistry 1997; 36:1035310363.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

136 de 138

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

557. Angles-Cano E, Rojas G. Apolipoprotein(a):


structure-function relationship at the lysine-binding site and
plasminogen activator cleavage site. Biol Chem
2002;383:9399.
558. Tsurupa G, Ho-Tin-Noe B, Angles-Cano E, et al.
Identification and characterization of novel lysine-independent
apolipoprotein(a)-binding sites in fibrin(ogen) C-domain. J Biol
Chem 2003;278:3715437159.
559. Palabrica TM, Liu AC, Aronovitz MJ, et al. Antifibrinolytic
activity of apolipoprotein(a) in vivo: human apolipoprotein(a)
transgenic mice are resistant to tissue plasminogen
activator-mediated thrombolysis. Nat Med 1995;1:256259.
560. Mooser V, Mancini FP, Bopp S, et al. Sequence
polymorphisms in the apo(a) gene associated with specific
levels of Lp(a) in plasma. Hum Mol Genet 1995;4:173181.
561. Dahln GH. Lipoprotein(a), atherosclerosis and
thrombosis. Prog Lipid Res 1991;30:189194.
562. Longenecker JC, Klag MJ, Marcovina SM, et al. Small
apolipoprotein(a) size predicts mortality in end-stage renal
disease: the CHOICE study. Circulation 2002;106:28122818.
563. Hervio L, Durlach V, Girard-Globa A, et al. Multiple binding
with identical linkage: a mechanism that explains the effect of
lipoprotein(a) on fibrinolysis. Biochemistry
1995;34:1335313358.
564. Angles-Cano E, de la Pena Diaz A, Loyau S. Inhibition of
fibrinolysis by lipoprotein(a). Ann N Y Acad Sci
2001;936:261275.

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

137 de 138

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

565. Falc C, Estells A, Dalmau J, et al. Influence of


lipoprotein (a) levels and isoforms on fibrinolytic activitystudy
in families with high lipoprotein (a) levels. Thromb Haemost
1998;79:818823.
566. Novokhatny VV, Jesmok GJ, Landskroner KA, et al. Locally
delivered plasmin: why should it be superior to plasminogen
activators for direct thrombolysis? Trends Pharmacol Sci
2004;25:7275.
567. Nilsson TK, Mellbring G. Impact of immediate acidification
of blood on measurement of plasma tissue plasminogen
activator (tPA) activity in surgical patients. Clin Chem
1989;35:1999.
568. Gaffney PJ, Curtis AD. A collaborative study to establish
the second international standard for tissue plasminogen
activator (tPA). Thromb Haemost 1987;59:10851087.
569. Grimaudo V, Hauert J, Bachmann F, et al. Diurnal variation
of the fibrinolytic system. Thromb Haemost 1988;59:495499.
570. Ridker PM, Vaughan DE, Stampfer MJ, et al. Endogenous
tissue-type plasminogen activator and risk of myocardial
infarction. Lancet 1993; 341:11651168.
571. Nordenhem A, Wiman B. Tissue plasminogen activator
(tPA) antigen in plasma: correlation with different tPA/inhibitor
complexes. Scand J Clin Lab Invest 1998;58:475483.
572. Rijken DC, Sakharov DV. Basic principles in thrombolysis:
regulatory role of plasminogen. Thromb Res 2001;103(Suppl.
1):S41S49.
573. Sakharov DV, Plow EF, Rijken DC. On the mechanism of

24/06/2006 01:14 p.m.

Ovid: Hemostasis and Thrombosis: Basic Principles and Clinical Practice

138 de 138

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

the antifibrinolytic activity of plasma carboxypeptidase B. J Biol


Chem 1997;272: 1447714482.

24/06/2006 01:14 p.m.

You might also like