Download factor xiii: a coagulation factor with multiple

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

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

Document related concepts

Complement component 4 wikipedia , lookup

Thrombophilia wikipedia , lookup

Plateletpheresis wikipedia , lookup

Hemolytic-uremic syndrome wikipedia , lookup

Platelet wikipedia , lookup

Transcript
Physiol Rev 91: 931–972, 2011
doi:10.1152/physrev.00016.2010
FACTOR XIII: A COAGULATION FACTOR WITH
MULTIPLE PLASMATIC AND CELLULAR
FUNCTIONS
László Muszbek, Zsuzsanna Bereczky, Zsuzsa Bagoly, István Komáromi, and Éva Katona
Clinical Research Center and Thrombosis, Haemostasis and Vascular Biology Research Group of the Hungarian
Academy of Sciences, University of Debrecen, Medical and Health Science Center, Debrecen, Hungary
L
I.
II.
III.
IV.
V.
VI.
VII.
VIII.
INTRODUCTION
GENERAL OVERVIEW ON STRUCTURE,...
ACTIVATION AND REGULATION OF...
THE ACTION OF ACTIVATED FACTOR XIII...
FACTOR XIII IN WOUND HEALING AND...
OTHER EFFECTS OF FACTOR XIII
FACTOR XIII AS AN INTRACELLULAR...
SUMMARY, CONCLUSIONS, AND...
931
932
940
945
947
952
957
961
became apparent only after Duckert et al. (94) demonstrated that the severe bleeding diathesis of a patient was
due to the deficiency of fibrin stabilizing factor. In 1963,
soon after this clinical finding, the International Committee
on Blood Clotting Factors acknowledged this protein as a
clotting factor and termed it factor XIII. In 2007 the International Society of Thrombosis and Haemostasis, Scientific
and Standardization Committee published a recommendation on the terms and abbreviations concerning FXIII (258),
which will be followed throughout this review.
I. INTRODUCTION
The history of blood coagulation factor XIII (FXIII) started
more than 70 years ago, when Barkan and Gaspar (33)
observed that fibrin clots formed in the presence of Ca2⫹
become insoluble in weak bases. Some 20 years later, Robbins (314) performed the experiment with purified fibrinogen and concluded that in addition to Ca2⫹, a “serum factor” is also needed to render the clot insoluble in weak acids
and bases. Laki and Lóránd (196) and Lóránd (221, 222)
were the first who realized that the serum factor, which
made the fibrin clot insoluble in concentrated urea solution,
was thermo-labile and nondialyzable, and they called this
protein fibrin stabilizing factor. Fibrin stabilizing factor was
then purified from plasma, and the enzymatic nature of it
was first revealed and characterized by Loewy and co-workers (215–219). The clinical significance of these findings
For a relatively long time, FXIII did not raise much interest
among scientists, and research on this field was practiced
only by a few devotees, although their contributions were
essential for later developments. It was revealed that
FXIII is a zymogen, and in contrast to all other zymogen
clotting factors, its active form (FXIIIa) is a transglutaminase (TG; protein-glutamine: amine ␥-glutamyltransferase, EC 2.3.2.13) that forms ⑀(␥-glutamyl)lysyl crosslinks between two polypeptide chains. The mechanism of
the activation of zymogen FXIII and the basic enzymology
of FXIIIa were clarified. FXIII was also found in platelets,
and structural differences between plasma FXIII (pFXIII)
and cellular factor (cFXIII) were revealed. Clinical data on
the consequences of inherited FXIII deficiency provided important pieces of information on features of the disease and
on the physiological role of FXIII in hemostasis. Neverthe-
931
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
Muszbek L, Bereczky Z, Bagoly Z, Komáromi I, Katona E. Factor XIII: A Coagulation
Factor With Multiple Plasmatic and Cellular Functions. Physiol Rev 91: 931–972,
2011; doi:10.1152/physrev.00016.2010.—Factor XIII (FXIII) is unique among clotting factors for a number of reasons: 1) it is a protransglutaminase, which becomes
activated in the last stage of coagulation; 2) it works on an insoluble substrate; 3) its
potentially active subunit is also present in the cytoplasm of platelets, monocytes, monocyte-derived
macrophages, dendritic cells, chondrocytes, osteoblasts, and osteocytes; and 4) in addition to its
contribution to hemostasis, it has multiple extra- and intracellular functions. This review gives a
general overview on the structure and activation of FXIII as well as on the biochemical function and
downregulation of activated FXIII with emphasis on new developments in the last decade. New
aspects of the traditional functions of FXIII, stabilization of fibrin clot, and protection of fibrin against
fibrinolysis are summarized. The role of FXIII in maintaining pregnancy, its contribution to the wound
healing process, and its proangiogenic function are reviewed in details. Special attention is given to
new, less explored, but promising fields of FXIII research that include inhibition of vascular permeability, cardioprotection, and its role in cartilage and bone development. FXIII is also considered as
an intracellular enzyme; a separate section is devoted to its intracellular activation, intracellular
action, and involvement in platelet, monocyte/macrophage, and dendritic cell functions.
MUSZBEK ET AL.
less, FXIII used to be considered a stepchild among clotting
factors for a number of reasons: 1) it falls outside the mainstream of the clotting cascade, and it only starts working
after the visible end result, fibrin formation, had been completed; 2) as opposed to other clotting enzymes, it is making
rather than breaking peptide bonds; and 3) it is working on
an insoluble substrate matrix that is rather unattractive for
enzymologists.
This review, besides giving a short overview on earlier discoveries concerning structure and biochemical functions of
FXIII, mainly deals with recent developments on the regulation of FXIII, with its role in fibrinolysis, and with its
functions which partially or completely lie outside hemostatic mechanisms. The latter topic includes the cellular effects of FXIII and its involvement in angiogenesis and
wound repair as well as in maintaining pregnancy. The
intracellular activation and function of FXIII will also be
discussed.
II. GENERAL OVERVIEW ON STRUCTURE,
GENE, POLYMORPHISMS, AND
EXPRESSION OF FACTOR XIII
The structure and synthesis of FXIII have been reviewed
earlier in detail (124, 263); here only a general overview
with emphasis on new discoveries will be provided. FXIII is
a pro-TG that circulates in plasma in tetrameric form
(FXIII-A2B2) (124, 259, 263). It consists of two potentially
active, catalytic A subunits (FXIII-A) and two carrier/inhibitory B subunits (FXIII-B). FXIII-B is in excess; in the
932
A. Factor XIII A Subunit and Cellular Factor
XIII
FXIII-A, a protransglutaminase, belongs to the family of
TGs. Eight members of the family (FXIII-A and TG1-TG7)
possess potential enzymatic activity; while an additional
protein, erythrocyte band 4.2, has a similar domain structure but, due to the replacement of active site amino acid,
lacks enzymatic activity. (A more detailed review on TGs is
provided in Reference 95.) All TGs are monomers; only
FXIII-A forms a dimeric structure. The three-dimensional
structure of three human TGs, FXIII-A2, TG-2, and TG-3,
has been resolved (14, 213, 377, 383). It is interesting that
despite the modest sequential agreement, their secondary
structure shows striking similarity. A detailed comparison
of the structure of the three TGs is given in Reference 187.
FXIII-A consists of 732 amino acids including an initiator
methionine (http://www.uniprot.org/uniprot/P00488); its
molecular mass is 83 kDa. Serine is in the penultimate position, which favors the removal of initiator methionine and
then N-acetylation. The NH2-terminal amino acid in the
mature molecule is, indeed, an acetylated serine residue. In
the following, amino acid numbering starting with the Ser
residue will be used. No hydrophobic leader sequence could
be identified in the primary structure of FXIII-A (144).
There are nine cysteine residues in the protein, including the
active site cysteine (Cys314), none of which forms disulfide
bonds. Although the molecule has potential sites for N-glycosylation, no carbohydrate residue is detected. FXIII-A is
expressed in all vertebrates investigated so far. A phylogenetic tree and sequence identity in different species are
shown on FIGURE 1. Recombinant FXIII-A2 (rFXIII) has
been expressed in Escherichia coli (45), in Saccharomyces
cerevisiae (42, 157), in Schizosacharomyces pombe (52),
and even in tobacco plant cells and whole tobacco plants
(108). Features of rFXIII are identical to those of cFXIII.
The first atomic resolution structure of FXIII-A2 was published (383) one and a half decades ago. Two crystal forms,
orthorhombic (383) and monoclinic (377), were obtained
for FXIII-A2, and X-ray crystallography of both crystal
forms resulted in essentially the same structure. Studies on
the X-ray structure of rFXIII-A2 (102, 377, 382, 383) revealed numerous structural details of FXIII-A2 and provided invaluable information on the structure-function relationships. The X-ray structure was also a great help in
understanding the structural consequences of mutations
causing FXIII deficiency (173).
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
Discoveries starting from the 1980s changed this trend, and
since then FXIII has become an increasingly attractive field
of research. The milestones of FXIII research were the following: 1) the realization that FXIII plays a highly important role in the regulation of fibrinolysis. 2) The discovery of
cFXIII in monocytes and different types of macrophages,
including tissue macrophages, revealed that this proenzyme
is present in most organs and tissues of the body. 3) The
primary protein structure and the genomic structure of
FXIII subunits were published in the second half of the
1980s followed by the exact identification of molecular genetic defects in an increasing number of FXIII deficient patients. In 1994 the three-dimensional structure of cellular
FXIII has also been published. 4) The availability of recombinant cFXIII (rFXIII) significantly promoted basic biochemical studies. 5) New, relatively simple reliable methods
for the determination of FXIII activity and concentration in
the plasma became available and made large-scale clinical
studies possible. 6) The discovery of FXIII polymorphisms
induced biochemical and clinical studies to explore their
consequences. 7) It was realized that FXIII is a multifunctional protein that, beside hemostasis, plays an important
role in a wide variety of physiological and pathological
processes.
plasma, ⬃50% of it exists in free noncomplexed form.
FXIII-A dimer (FXIII-A2), but not FXIII-B, is also present in
the cytoplasm of certain cells, particularly of platelets and
monocytes/macrophages. The primary structure of both
subunits has been determined by cDNA cloning and amino
acid sequence analysis (127, 145, 146, 359).
MULTIPLE FUNCTIONS OF FACTOR XIII
FXIII-A consists of four main structural domains, ␤-sandwich (amino acid 38 –184), catalytic core (185–515), ␤-barrel 1 (516 – 628), and ␤-barrel 2 (629 –731) domains, plus
an NH2-terminal activation peptide (AP-FXIII) (FIGURE 2).
The latter is cut off by thrombin during activation (see sect.
IIIA for details). The sandwich and barrel domains almost
exclusively consist of ␤-sheets with only a few small helical
structural elements in the sandwich and ␤-barrel 1 domains.
The core domain contains both ␤-sheets and helices. On the
basis of thermal denaturation studies (92, 193), the core
domain can be further divided into NH2- (189 –332) and
COOH-terminal (333–515) subdomains. Two nonproline
cis peptide bonds were observed in the core domain (377)
that also exists in human TG-2 (213) and TG-3 (14). The
FXIII-A subunit monomers are arranged in a dimeric structure resembling a hexagon with C2 symmetry, in which the
two central core domains are surrounded by the six ␤-sheet
domains (FIGURE 2).
The X-ray studies also revealed that the arrangement of the
amino acid residues playing a key role in the catalytic process (Cys314, His373, Asp396) resembles the catalytic triad
in cysteine proteases (FIGURE 3A). Therefore, it was assumed that the catalytic reaction corresponds to a reverse
proteolysis (294). However, due to the active site specificity
of the enzyme, it catalyzes interchain peptide bond formation between the glutamine and lysine side chains instead of
peptide chain elongation. In the reaction catalyzed by cys-
teine proteases, the stabilization of the intermediates by the
so-called oxyanion hole is of crucial importance. Based on
the structural analogy with papain, it has been assumed that
the NH group of the indole side chain of Trp279 and the
NH group of Cys314 backbone form the oxyanion hole
(294) (FIGURE 3A). Trp279 is highly conserved in vertebrate
TGs, and its direct involvement in the TG reaction has been
shown by kinetic study with TG-2 (147).
Although the geometrical arrangement of the catalytic triad
and the oxyanion hole in the catalytic domain corresponds
to the arrangements in cysteine proteases (288), X-ray
structure also revealed that the catalytic cysteine residue is
completely buried by AP-FXIII (FIGURE 3A). A strong saltbridge between the side chain of Arg11= in AP-FXIII extending over from the opposite subunit and the side chain of
Asp343 of the core domain enforces AP-FXIII into a position
that makes Cys314 inaccessible for the substrate and even for
the solvent molecules. Therefore, the removal or disposition of
AP-FXIII is a prerequisite of activation. However, this is not
sufficient since the accessibility of the active site to the substrate is also prevented by the strong hydrogen bond that exists
between the O␩ atom of Tyr560 and S␥ atom of Cys314
residues (FIGURE 3A). Tyr560 is located on a loop of the ␤-barrel 1 domain, which means that the transposition of this loop
from the immediate vicinity of the active site is also required
for the interaction with the substrate. Considering the extreme
stability of barrel domains (193) and the size of the natural
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
933
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
FIGURE 1 Phylogenetic tree and sequence
identity of A subunit of factor XIII (FXIII-A) in
different species. A: phylogenetic tree of
FXIII-A was constructed by the BIONJ version
(109) of neighbor-joining algorithm using the
SplitsTree4 software (141) downloaded from
the www.splitstree.org website. FXIII-A protein sequences were aligned by ClustalX software (198). Distance scale represents the
number of differences between sequences
in species since their divergence; 0.1 means
10% difference between two sequences. B:
identity percent represents the degree of homology in different organisms compared with
human FXIII-A protein and coding DNA. Data
were obtained from the HomoloGene database (www.ncbi.nlm.nih.gov/sites/homologene). The amino acid sequence of salmon
FXIII-A was downloaded from the Universal
Protein Resource (UniProt) web server
(http://www.uniprot.org), and its identity
percent value was calculated by BLAST software (www.blast.ncbi.nlm.nih.gov) using
the alignment obtained by ClustalX software
(198).
MUSZBEK ET AL.
FIGURE 2 Ribbon and tube representation of the zymogen
rFXIII-A2 homodimer based on its X-ray structure (377) and visualized by the Chimera software (299). The relative positions of the
constituting domains are marked for only one of its monomeric
FXIII-A subunit. Helical and ␤-sheet structural elements are colored
orange and purple, respectively. The disordered (random coil like)
structural elements and loops are shown in cyan, and the activation
peptides are depicted in bright green.
substrates, the removal of the whole ␤-barrel 1 from its original position is an even more likely option.
Ca2⫹ binding sites on the surface of FXIII-A2 are essential for
the activation of FXIII. FXIII-A, but not FXIII-B, tightly binds
The gene coding for human FXIII-A (F13A1) spans over 160
kb and has been localized to chromosome 6p24 –25 (46). It is
transcribed into a 3.9-kb mRNA, with an 84-bp 5=-untranslated region, a 2.2-kb open reading frame, and a 1.6-kb 3=untranslated region. F13A1 contains 15 exons and 14 introns
(144). Exon I consists of the 5= noncoding region, and exon II
encodes AP-FXIII (FIGURE 4). The ␤-sandwich domain, the
catalytic core domain, and the two ␤-barrel domains are encoded by exons II-IV, exons IV-XII, exons XII-XIII, and exons
XIII-XV, respectively.
shows the polymorphisms and their location in
the coding region of FXIII-A gene. Data provided by the
International HapMap project (www.hapmap.org) on the
frequency of the polymorphisms in the Caucasian, African,
FIGURE 4
FIGURE 3 Structures around the catalytic triad (A) and the main Ca2⫹ binding site (B) of FXIII-A. Structural
elements are shown by “ribbon and tube” representation; helical and ␤-sheet structural elements are colored
orange and purple, while disordered (random coil like) structural elements and loops are shown in light green.
Side chains of selected amino acid residues are represented by balls and sticks. The N, C, O, and S atoms are
colored blue, gray, red, and yellow, respectively. A: the amino acids of the catalytic triade (Cys314-His373Asp396) and Trp279 participating in the formation of oxyanion hole are shown together with interacting amino
acid residues that fix the structure in inactive conformation and bury the active site cysteine. The X-ray
structure published by Weiss et al. (377) was used for the preparation of the picture (377). Arg11= represents
the Arg residue in AP-FXIII of the opposite subunit. B: amino acid residues implicated in the binding of Ca2⫹
(green ball) at the main Ca2⫹ binding site. The structure based on the X-ray coordinates published by Fox et al.
(102) was visualized by the Chimera software (299).
934
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
one Ca atom with a Kd of 0.1 mM as determined by equilibrium dialysis and fluorescence studies (209), while a Kd of 0.51
mM was determined by 43Ca NMR experiments (19). These
reports also suggest the existence of additional low-affinity
Ca2⫹ binding sites. From X-ray structural studies of FXIII-A2
cocrystallized with Ca2⫹, Str2⫹, and Yb3⫹, a single main cation binding site per subunit was identified (102). The carboxylate group of Asp438, Glu485, and Glu490 side chains, the
carbonyl O atom of Asn436, as well as the backbone carbonyl
O atom of the Ala457 residue participate in cation binding
(either directly or through a water bridge) to a different extent
depending on the nature of the cation (FIGURE 3B). The backbone carbonyl O atom of Ala457 and the carboxylate groups
of Glu485 and Glu490 side chains have been reported to be
directly involved in Ca2⫹ binding (102, 195).
MULTIPLE FUNCTIONS OF FACTOR XIII
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
FIGURE 4 Distribution of SNPs published so far within the coding region of F13A1 gene. The “A” of the
translation initiation codon “ATG” is considered as nucleotide one, and nucleotide positions within cDNA are
given accordingly. The amino acid numbering is given according to Ichinose et al. (144). Exons encoding a
particular domain are color-coded. The mutations resulting in amino acid replacements are depicted in purple,
while silent mutations are shown in black.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
935
MUSZBEK ET AL.
tive effect against venous thromboembolism (378) and
myocardial infarction (370). The complex interrelationship of FXIII polymorphisms with the risk of thrombotic
disease is beyond the scope of this review; interested
readers should consult a most recent review (259). The
biochemistry of other common FXIII-A polymorphisms
has not been investigated in such details as that of
Val34Leu polymorphism. The Phe204 allele of FXIII-A
Tyr204Phe polymorphism was reported to be associated
with decreased pFXIII level and activity, whereas the
Leu564 allele of the Pro564Leu variant resulted in lower
FXIII plasma level with increased FXIII activity (22,
107). However, these results still need to be confirmed.
FXIII-A is expressed primarily in cells of bone marrow origin. It is present in platelets in huge quantity (55, 228). The
average FXIII-A content of a single platelet was estimated to
be 60 ⫾ 10 fg, which corresponds to 3% of total platelet
proteins (176). The concentration of FXIII-A is ⬃100- to
150-fold higher in platelet cytoplasm than in plasma.
FXIII-A is also present in megakaryocytes (10, 179, 231,
256) and in their precursor cells (10). Megakaryocytes synthesize FXIII-A and package it, together with its encoding
Table 1 Racial variations of the polymorphisms in factor XIII-A and factor XIII-B subunit genes
Nucleotide Variation
F13A1
c.103G⬎T
c.614A⬎T
c.996A⬎C
c.1652C⬎T
c.1694C⬎T
c.1704A⬎G
c.1696T⬎A
c.1951G⬎A
c.1954G⬎C
F13B
c.344G⬎A
c.456G⬎A
c.765C⬎T
c.1049A⬎G
c.1707T⬎G
c.1806T⬎C
c.1952 ⫹ 144 C⬎G
Allele Frequency
Amino Acid
Variation
CEU
YRI
JPT
HCB
p.V34L
p.Y204F
p.P331P
p.T550I
p.P564L
p.E567E
p.L588Q
p.V650I
p.E651Q
0.767/0.233
0.983/0.017
0.853/0.147
1
0.758/0.242
0.890/0.110*
0.975/0.025
0.950/0.050
0.775/0.225
0.883/0.117
1
0.933/0.067
0.992/0.008
0.864/0.136
0.991/0.009
1
0.942/0.058
0.742/0.258
1
1
1
1
0.705/0.295
1
1
0.909/0.091
0.909/0.091
1
1
0.989/0.011
1
0.733/0.267
1
1
0.911/0.089
0.911/0.089
p.R95H
p.T132T
p.C235C
p.H330R
p.D549E
p.N582N
-
0.075/0.925
0.925/0.075
1
1
1
0.492/0.508
0.85/0.142
0.725/0.275
0.307/0.693
0.983/0.017
0.933/0.067
0.908/0.092
0.842/0.158
0.98/0.017
0.034/0.966
1
1
0.989/0.011
0.966/0.034
0.244/0.756
0.42/0.578
0.044/0.956
1
1
0.989/0.011
0.956/0.044
0.189/0.811
0.35/0.644
In the case of F13A1 and F13B polymorphisms, amino acid positions are numbered according to Ichinose et
al. (144) and according to Bottenus et al. (50), respectively. Allele frequency data of the International HapMap
project (www.hapmap.org) were used; the first numbers represent the frequency of ancestral allele. CEU, Utah
residents with ancestry from Northern and Western Europe; YRI, African population in Yoruba in Ibadan,
Nigera; JPT, Japanese in Tokyo, Japan; HCB, Han Chinese in Beijing, China. *As in the case of c.1704A⬎G
mutation, no HapMap data for CEU were available; data were taken from the Centre d’Etude du Polymorphisme
Human (CEPH) project (www.cephb.fr).
936
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
and Asian populations demonstrate considerable racial
variation (TABLE 1). Detailed studies on the biochemical
and clinical consequences of the polymorphisms were carried out only in the case of Val34Leu polymorphism, a
primarily Caucasian mutation first described by Mikkola et
al. (244). This mutation increases the rate of FXIII activation (23, 31, 374) (see sect. IIIA for details). The structure of
fibrin clots is also influenced by FXIII-A Val34Leu polymorphism (23), and this effect is modulated by fibrinogen
concentration (211). At high fibrinogen levels, plasma samples from homozygotes for the Leu34 allele form clots having looser structure, thicker fibers, and increased permeability, while at low fibrinogen concentrations fibrin
meshwork had thinner, more tightly packed fibers, and
lower permeability. Practically no fibrinogen concentration-dependent changes were observed in the plasma
samples of wild-type individuals. The effect of the
Val34Leu polymorphism on the specific activity of
FXIIIa was controversial for a while; it is now clear that
the specific activity of fully activated pFXIII (23, 31),
cFXIII (374), and rFXIII-A2 (201) of different FXIII-A
Val34Leu genotypes are identical. Although there are
contradictory results on the clinical implications of this
polymorphism, meta-analyses demonstrated its protec-
MULTIPLE FUNCTIONS OF FACTOR XIII
mRNA, into newly formed platelets (5, 6, 231). It is now
clear that FXIII-A in platelets is essentially of cytoplasmic
localization (51, 220, 278, 346). Platelets could take up a
tiny amount of FXIII-A2B2 from the plasma which appears
in ␣-granules (233); this amount is negligible compared
with the amount of FXIII-A in the cytoplasm.
The presence of FXIII-A mRNA and the synthesis of
FXIII-A have also been verified in the above cell types (4, 5,
190, 271). The transcriptional regulation of cell type-specific expression of FXIII-A was investigated in monocytoid
(U937) and megakaryocytoid (MEG-01) cell lines containing FXIII-A mRNA (178). Reporter gene assays revealed that
a 5=-fragment was sufficient to support the basal expression in
the above cells, but not in other cell types. Promoter elements
for a myeloid-enriched transcription factor (MZF-1) and for
the ubiquitous transcription factors, SP-1 and NF-1, seem to
be important for the basal FXIII-A expression. Another region
had enhancer activity in MEG-01 cells but silencer activity in
B. Factor XIII B Subunit
FXIII-B is a glycoprotein consisting of 641 amino acids
and containing 8.5% carbohydrate; its molecular mass is
⬃80 kDa. It is a typical mosaic protein consisting of 10
short tandem repeats, called sushi domains or GP-I structures, each containing ⬃60 amino acids and held together by a pair of internal disulfide bonds (FIGURE 5).
The same structures have also been found in more than
20 other proteins, including ␤2-glycoprotein I, many of
which are encoded by genes clustered in chromosome 1
band q32 (143). By electron microscopy, the B subunit
appears as a thin, flexible, and kinked strand (58).
FXIII-B has not been crystallized, and no reliable information on its three-dimensional structure is available.
While FXIII-A in the absence of FXIII-B forms homodimer, contradictory publications concerning the
state of noncomplexed FXIII-B were obtained. Gel filtration of purified FXIII-B suggested the presence of FXIII-B
dimers (335), while sedimentation analysis revealed monomeric form (58). In a most recent report, gel filtration
of recombinant FXIII-B supported the former finding
(350). There is no good explanation for the discrepancy,
and in plasma conditions, the state of free FXIII-B remains to be explored.
The gene of FXIII-B subunit (F13B) is located at position
1q31–32.1; it is ⬃28 kb in length and composed of 12
exons producing a 2.2 kb mRNA (146). Exons are interrupted by 11 introns. Exon I encodes a 20-amino acid
leader sequence characteristic for proteins that are secreted
via the classical secretory pathway (50, 169, 375). Exon
II-XI code for the sushi domains, each of which is encoded
by a single exon. The sushi domains show a high degree of
homology, suggesting gene duplication and exon shuffling
during evolution (292, 293). The last exon codes for a
COOH-terminal region of FXIII-B, for the 3=-untranslated
region, and for the polyA tail. F13B is directly regulated by
transcription factor HNF1␣ and HNF4␣ (143, 214). Accordingly, FXIII-B is expressed in the liver and secreted by
hepatocytes (146, 149, 266).
The polymorphic nature of FXIII-B was revealed a long
time ago (43, 81, 207). On the basis of isoelectric focusing experiments, three major population-associated phenotypes were described: FXIII-B*1, FXIII-B*2, and
FXIII-B*3, characteristic of European, African, and
Asian populations, respectively. Since the first report of a
F13B gene polymorphism by Board (43), several new
polymorphisms have been described. FXIII-B polymorphisms and their frequency in different races are summarized on FIGURE 6 and in TABLE 1. By screening all the
exons of F13B, Komanasin et al. (186) found three poly-
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
937
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
FXIII-A was also detected in monocytes (257) and in their
bone marrow precursor cells (10). Like platelets, monocytes
also lack FXIII-B. A most recent study revealed that FXIII-A
may be differently expressed in monocyte subsets; higher
FXIII-A mRNA expression was demonstrated in Ly-6Chigh
monocytes than in their Ly-6Clow counterparts (358).
FXIII-A in monocytes is also of cytoplasmic localization,
although during differentiation into macrophages it also
appears in the nucleus (7). The results of the first discovery
were soon confirmed, and FXIII-A was also detected in a
number of monocyte-derived macrophages including macrophages of different serous cavities, alveolar macrophages,
tumor-associated macrophages, histiocytic and dendritic
reticulum cells of lymph nodes, connective tissue histiocytes, perivascular dendritic macrophages, dermal dendrocytes etc. (reviewed in Refs. 6, 263). FXIII-A has also been
demonstrated in chondrocytes, osteoblasts, and osteocytes
(279) (see sect. VID). In early studies, FXIII-A was also
identified in placenta and in uterus (48, 61). With the use of
monocyte and macrophage differentiation markers, it
has been clearly demonstrated that FXIII-A is confined to
monocyte-derived tissue macrophages in the placenta (9,
171) and in the uterus (11) as well. The expression of
FXIII-A in monocytes is retained after malignant transformation; in fact, its expression in the respective cells is
upregulated in acute myelomonocytic, monocytic leukemias and in chronic myelomonocytic leukemias (172). It
turned out that the detection of FXIII-A by flow cytometry in leukemic cells is a useful intracellular marker in
the classification of acute myeloid leukemias. It is interesting that FXIII-A becomes expressed in part of the lymphoblasts in acute B cell lymphoid leukemia, while it is
absent in lymphocytes, normal lymphoid precursors, and
mature lymphocytes from B cell chronic lymphoid leukemia (183).
U937 cells, and the GATA-1 element was found to be responsible for the enhancer activity.
MUSZBEK ET AL.
morphisms in healthy Caucasian subjects. An A to G
transversion within exon III (rs6003) leads to His to Arg
amino acid exchange at codon 95 in the second sushi
domain of the mature protein. The frequency of FXIII-B
Arg95 carriers among healthy Caucasians is ⬃15%, it is
more frequent among Africans (also named as FXIII-B*2
at the phenotypic level), and it is missing from the Asian
population (155, 186). The polymorphism did not influence
FXIII-A, FXIII-B, or pFXIII antigen levels. Increased subunit
dissociation was found in plasma from subjects possessing the
Arg allele. However, when the variants were purified to homogeneity and binding was analyzed by steady-state kinetics,
no difference was observed (186).
Most recently, a C-to-G change at position 29756 in
intron K (IVS11⫹144) leading to a novel splice acceptor
site was described (155, 321). This polymorphism results
in allele-specific splicing products, and a protein 15
amino acids longer at the COOH terminus than its wildtype counterpart is synthesized. The variant sequence
includes two additional lysine and one glutamic acid res-
938
idues. These charged amino acids change the pI of the
protein. The polymorphism characteristically occurs in
Asians; it corresponds to FXIII-B*3 at the phenotype
level. Its frequency is less in Caucasian populations, and
it does not seem to be present among Africans. Although
such a profound structural change would be expected to
alter some of the biochemical features of the molecule,
this possibility has not been explored, yet.
C. Factor XIII Complex in Plasma
The origin of FXIII-A in pFXIII complex has been a debate
for a long period of time. Studies characterizing phenotypes
of plasma FXIII subunit A after bone marrow and liver
transplantation clearly indicated that the major source of
plasma FXIII-A is a cell population or cell populations of
bone marrow origin (303, 379). In one of the studies, a
minor source of recipient FXIII-A, independent of donor
hematopoesis, was also detected (303); in theory, such
source(s) could be long-persisting tissue macrophages or
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
FIGURE 5 Schematic structure of FXIII-B consisting of 10 sushi domains each held together by two pairs of
disulfide bonds. Amino acids in the leader sequence are depicted in purple, and those in the mature protein are
shown in cyan. The solid stars point to the site of N-glycosylation, and the empty star points to a potential
glycosylation site to which no carbohydrate is attached.
MULTIPLE FUNCTIONS OF FACTOR XIII
liver cells. However, FXIII-A and its encoding mRNA could
be detected in hepatocytes only in trace amounts using
highly sensitive techniques (5, 6), and there is no evidence
on the contribution of tissue macrophages to FXIII-A in the
plasma. Bone marrow ablation in patients undergoing autologous peripheral blood stem cell transplantation decreased platelet count to a very low level (⬎90% decrease),
while the reduction of plasma FXIII level was only ⬃25%
(152). It has also been shown that in long-standing severe
thrombocytopenia, the reduction of plasma FXIII was less
than expected (179). These findings suggest that in the case
of highly impaired production of FXIII-A by bone marrow
cells, synthesis of FXIII-A could be taken over by other cell
types. It is to be noted that in early embryonic life, well
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
939
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
FIGURE 6 Distribution of SNPs published
so far within the coding region of F13B
gene. The “A” of the translation initiation
codon “ATG” is considered as nucleotide
one, and nucleotide positions within cDNA
are given accordingly. The amino acid numbering is given according to Bottenus et al.
(50). The mutations resulting in amino acid
replacements are depicted in purple, and
silent mutations are in black. The intronic
mutation resulting in an alternative splice
site is shown in green.
MUSZBEK ET AL.
Only the liver has been demonstrated as the source of
FXIII-B in plasma (see above). This conclusion was also
supported by phenotyping plasma FXIII-B following liver
transplantation. In this case, the phenotype of recipient’s
FXIII-B changed to the donor’s phenotype, while the phenotype of FXIII-A remained unchanged (379).
FXIII subunits derived from separate cellular sources form
tetrameric complex in the circulating blood. A reference
interval of 14 –28 mg/l has been established for FXIII-A2B2
in human plasma (175). It is surprising that there is still
uncertainty concerning the affinity of the two types of FXIII
subunits. Apparent binding constants of 4 ⫻ 10⫺7 M (305)
and 8.06 ⫻ 10⫺8 M (186) were reported. However, when
calculating with these binding constants, a significant part
of FXIII-A should be in free form, which is evidently not the
case. In our laboratory one magnitude lower binding constant (2.73 ⫻ 10⫺9 M) was measured; this value fits better
with the fact that FXIII-A is present in the plasma in practically fully complexed form. In normal conditions, FXIII-B
is in excess in the plasma, and ⬃50% of it circulates in free,
noncomplexed form (384). In patients with FXIII-A deficiency, the total amount of FXIII-B in the plasma is reduced
while the concentration of free B subunits remains constant.
The FXIII-A domain(s) that are responsible for the association with FXIII-B have not been identified. Most recently,
binding analysis with various truncated forms of rFXIII-B
suggested that the first sushi domain was responsible for the
binding of FXIII-B to FXIII-A (350).
In the plasma, practically all FXIII molecules are bound to
fibrinogen (Kd ⬃10⫺8 M) and the association is independent of the presence of calcium ions (123, 125, 215). The
␥=-chain in ␥A␥= fibrinogen (fibrinogen 2) provides the major binding site for FXIII (91, 250). The ␥=-chain is present
in ⬃15% of plasma fibrinogen. Compared with the ␥A
chain, the variant chain contains a 20-amino acid exten-
940
sion, which is the binding site for FXIII through its B subunit and also for thrombin exosite 2.
III. ACTIVATION AND REGULATION OF
FACTOR XIII: THE BIOCHEMICAL
ACTION OF ITS ACTIVE FORM
A. Activation of Factor XIII in the Plasma
and Within Cells
The activation of pFXIII occurs in the final phase of the
clotting cascade by the concerted action of thrombin and
Ca2⫹ (FIGURE 7). In the initial step of the reaction, thrombin
cleaves off AP-FXIII from the NH2 terminus of FXIII-A by
hydrolyzing the Arg37-Gly38 peptide bond. Then, in the
presence of Ca2⫹, the inhibitory B subunits dissociate,
which is a prerequisite for the truncated FXIII-A dimer
(FXIII-A2=) to assume an enzymatically active conformation (FXIII-A2*). The conformational change of FXIII-A2=
resulting in an active TG also requires Ca2⫹. The proteolytic cleavage of FXIII-A by thrombin considerably weakens the interaction between A and B subunits (224, 305).
The binding of Ca ions to the high-affinity Ca2⫹ binding
sites on the A subunits is sufficient to dissociate the subunits
and to activate the released A= dimer (138, 210). Interestingly, in the presence of Ca2⫹, complete activation of the
FXIII-A dimer occurs if only one AP-FXIII is released from
one of the two A subunits: in other words, the dimer of a
cleaved and an uncleaved A subunit (FXIII-A*A°) possesses
full enzymatic activity (138). It is to be noted that Siebenlist
et al. (342) reported a low innate activity of the zymogen
FXIII-A2B2 that can slowly cross-link fibrin(ogen) in the presence of Ca2⫹, but it is ineffective on low-molecular-weight
substrates (342). It is not clear if such an activity is due to the
presence of a small fraction of free FXIII-A2 that remains uncomplexed or dissociates from the complex and undergoes
nonproteolytic activation (see below) or it is the intrinsic property of the fibrin(ogen)-bound tetrameric zymogen.
Thrombin is not the only serine protease that could cleave
and activate FXIII in the presence of Ca2⫹. Several other
proteases, including batroxobin marajoensis (371), thrombocytin (273), trypsin (188, 334), and activated factor X
(FXa) (236) have been reported to be able to activate FXIII.
Although the cleavage site of these enzymes on FXIII-A has
not been identified, based on their substrate specificity and
on the Mr of the truncated FXIII-A, it is assumed that they
cleave the same Arg37-Gly38 peptide bond as thrombin,
and the active form of FXIII-A produced by these proteases,
like the thrombin activated form, also has an NH2-terminal
glycine. In contrast to thrombin, the proteolytic cleavage of
FXIII-A by FXa requires Ca2⫹. No evidence has been provided on the physiological implication of FXa-induced
FXIII activation. The mannan-binding lectin associated serine protease 1 (MASP1) that is involved in the complement
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
before the development of bone marrow, FXIII-A is present
in mesenchymal histiocytes, and by week 20, FXIII-A also
appears in liver cells (170). In summary, in normal conditions, FXIII-A present in pFXIII is predominantly synthesized by cells of bone marrow origin; however, in the case of
impaired bone marrow function, other not yet identified cell
types might take over the production of FXIII-A. Another
unresolved question is how FXIII-A is released from cells
that synthesize it. In the absence of leader sequence, the
classical secretory pathway does not operate. In a most recent paper, cFXIII in macrophages was found in association
with podosomes and other structures adjacent to the plasma
membrane, and FXIII-A was present in intracellular vesicles
positive for Golgi matrix protein 130, which has been implicated in the delivery of nonclassically secreted proteins to the
plasma membrane (75). Although this finding raises the possibility of a nonclassical secretory pathway, clear-cut evidence
supporting this hypothesis is still to be provided.
MULTIPLE FUNCTIONS OF FACTOR XIII
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
FIGURE 7 Different mechanisms of FXIII activation. FXIII-A=, A subunit of FXIII from which the activation
peptide was cleaved off; FXIII-A*, FXIII-A= transformed into an active transglutaminase in the presence of Ca2⫹;
FXIII-A°, noncleaved FXIII-A transformed into an active transglutaminase by the nonproteolytic mechanism in the
presence of Ca2⫹. Green and orange cylinders represent ␤-barrel and ␤-sandwich domains of FXIII-A, respectively. The central core domains in FXIII-A are depicted as horseshoes in magenta. The activation peptides are
shown as red loops. The elongated bended structure consisting of 10 pearls surrounding FXIII-A2 corresponds
to FXIII-B.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
941
MUSZBEK ET AL.
system and possesses thrombin-like activity also cleaves
FXIII-A at Arg37-Gly38, although at a much slower rate
than thrombin (191). It is not likely that FXIII activation by
MASP1 has any physiopathological importance. Recently,
it has been shown that limited cleavage of FXIII by human
neutrophil elastase resulted in the activation of both pFXIII
and cFXIII with TG activities of ⬃50% of thrombin-activated FXIII (28). MALDI-TOF analysis of the cleaved fragments and NH2-terminal Edman degradation of the truncated protein identified Val39-Asn40 as the primary cleavage site, and the NH2-terminal amino acid of this novel
active form of FXIII is Asn.
The presence of polymerizing fibrin enhances the activation
of pFXIII by ⬃100-fold (122, 123, 159, 210, 272). The
binding to fibrin makes the orientation of both pFXIII and
thrombin favorable for the proteolysis of FXIII-A at Arg37Gly38. Both fibrin I, which is devoid of fibrinopeptide A,
and fibrin II, which is devoid of fibrinopeptides A and B, are
effective accelerators. This effect is partly attributed to the
FXIII-B-driven optimal orientation of the tetramer molecule. It facilitates the interaction with thrombin and accelerates the thrombin-induced cleavage of FXIII-A in the
presence of fibrin (122, 137). In the absence of FXIII-B, i.e.,
in the case of cFXIII, no fibrin-induced promotion of the
removal of AP-FXIII could be observed (121, 137). Experiments with thrombin mutants demonstrated that His66,
Tyr71, and Asn74, surface residues of thrombin, are involved in high-specificity fibrin-enhanced factor XIII activation. These residues, representing a distinct interaction
site on thrombin (within exosite 1), are also employed by
thrombomodulin in the cofactor-enhanced activation of
protein C (300). Enhancement of factor XIII activation by
fibrin was competitively inhibited by thrombomodulin.
When the degree of fibrin ␥-chain cross-linking exceeds
942
Results concerning the effect of FXIII binding to fibrinogen ␥=-chain on FXIII activation are contradictory.
Moaddel et al. (247) reported that ␥A␥A fibrin accelerated the activation of pFXIII less efficiently than ␥A␥=
fibrin. In contrast, Siebenlist et al. (342) demonstrated
more rapid FXIII activation in the presence of ␥A␥A
fibrin(ogen) than with ␥A␥= fibrin(ogen). Atroxin, a
thrombin-like snake venom enzyme that does not bind to
the ␥= sequence, activated FXIII in the presence of ␥A␥A
and ␥A␥= fibrin(ogen) to the same extent. This finding
indicates that rate differences observed with thrombin
are related to the binding of thrombin rather than to the
binding of FXIII to the ␥= sequence.
In the plasma FXIII becomes effectively activated only on
the surface of newly formed fibrin, and the truncated active
dimer remains associated with its substrate; no thrombincleaved FXIII could be detected in the serum (341), while
following dissociation, FXIII-B becomes detached from fibrin. Fluorescence studies showed that cFXIII also binds to
fibrin, but with less affinity than pFXIII. The affinity, however, is increased upon cFXIII activation (137). Fibrin(ogen) also regulates the Ca2⫹ requirement for the activation
of truncated FXIII-A2= (78, 79). It brings down the optimal
Ca2⫹ concentration from ⬎10 mM to the physiological
interval present in the plasma.
Considering that the location of FXIII-A Val34Leu polymorphism is just 3 amino acids upstream from the thrombin
cleavage site, one would expect its influence on thrombininduced FXIII activation. Indeed, it was demonstrated with
both cFXIII (374) and pFXIII (23, 31) that the thrombininduced release of AP-FXIII from the Leu34 FXIII-A variant, as well as the consequent activation of FXIII, proceed
at a 2.5-fold higher rate than in the case of the Val34 variant. Faster activation of FXIII results in accelerated fibrin
cross-linking and in a higher rate of ␣2-PI incorporation
into fibrin (23, 31, 330, 374). The introduction of
Val34Leu and Val35Thr double mutations into FXIII-A
increased its rate of activation even further (by 7.6-fold)
(20). Investigation of the effect of FXIII-A Val34Leu genotype on thrombin-induced FXIII activation in the more
complex environment of human plasma led to a similar
conclusion (341). In plasma, the onset of FXIII activation
was initiated by fibrin polymerization independently of the
polymorphism; however, after initiation, the release of
Leu34 AP-FXIII occurred significantly faster than that of
Val34 AP-FXIII.
A few pieces of biochemical evidence suggest major conformational changes of FXIII during the process of activation;
however, the active form of FXIII has not been crystallized
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
The question whether AP-FXIII is released into plasma
after thrombin cleavage or it remains noncovalently associated with the truncated FXIII molecule has been addressed by X-ray crystallography and by biochemical
techniques. In the crystal prepared from FXIII-A2 treated
with thrombin and Ca2⫹, the AP-FXIII appeared in the
same position as it was in the zymogen (382). It is also
known that in certain conditions the proteolytic removal
of AP-FXIII is not a prerequisite of the active configuration (79, 223, 302). However, as discussed earlier, the
data obtained by X-ray crystallography do not fit well
with the biochemical data obtained with FXIIIa in solution. Recent findings obtained with antibodies specific
for free AP-FXIII indicate that in plasma AP-FXIII dissociates from the truncated parent molecule and appears in
the serum (287, 332). While noncleaved AP-FXIII is
strongly hydrogen bonded to the adjacent ␤-sandwich
and to the ␤-barrel 1 domain of opposite FXIII-A monomer, according to NMR and circular dichroism studies
its released form seems unstructured (287).
⬃40%, the promoter effect of fibrin is lost (210), i.e., the
formation of cross-linked product downregulates the activation of cross-linking enzyme.
MULTIPLE FUNCTIONS OF FACTOR XIII
While X-ray crystallographic study could not reveal significant difference in the structure of nonactivated and thrombin and Ca2⫹-activated FXIII-A2 (382), the above findings
suggest a considerable conformational change during the
transformation of FXIII-A2 into FXIII-A2*. It is interesting
that Pinkas et al. (301) trapped TG-2 in complex with an
inhibitor peptide, Ac-P(DON)LPF, that mimics a natural
Gln substrate and solved the X-ray crystal structure (301).
They demonstrated that the substrate analog stabilizes active TG-2 in an extended conformation, dramatically different from earlier TG structures. No such crystal structure
of FXIIIa has been produced. Based on the homology with
TG-2, a molecular model of FXIII-A2* with extended conformation has been developed, and this model showed good
correlation with the biochemical data (187).
Although in physiological conditions pFXIII is generally
activated by the concerted action of a protease (thrombin)
and Ca2⫹, proteolytic cleavage is not an absolute requirement for the activation of the tetramer molecule. In nonphysiological conditions, i.e., at extremely high Ca2⫹ concentrations (ⱖ100 mM), the A2B2 complex dissociates and
the nontruncated FXIII-A2 becomes transformed into an
active TG (FXIII-A2°) (79, 223). The occupancy of lowaffinity Ca2⫹ binding sites might play a role in this process.
In sharp contrast to the nonproteolytic activation of pFXIII,
low Ca2⫹ concentration is sufficient to induce a slow progressive activation of cFXIII at physiological ionic strength
(302). The rate of this nonproteolytic activation of cFXIII is
greatly enhanced by increasing the ionic strength, and complex formation with FXIII-B fully abrogates the activation
process (302). This finding suggests that in the case of
pFXIII, high Ca2⫹ concentration is only required for the
dissociation of FXIII-B from FXIII-A2.
In extracellular conditions, cFXIII could be activated by
thrombin and Ca2⫹ the same way as pFXIII, excluding the
dissociation of FXIII-B. It was suggested that cFXIII on
platelets is activated by the cysteine protease calpain (21).
However, in the intracellular environment, cFXIII does not
need proteolytic cleavage for activation, and during platelet
activation, no proteolytic truncation of FXIII-A occurs
(261, 262). The elevation of intracellular Ca2⫹ concentration, as a consequence of cell activation, seems sufficient to
bring about the enzymatically active configuration (FXIIIA2°). The activation of cFXIII by the nonproteolytic mechanism (i.e., by the elevation of intracellular Ca2⫹ concentration) represents the physiological way of cFXIII activation in platelets (261, 262) and most likely in monocytes (1)
as well. In platelets, elevation of intracellular Ca2⫹ concentration, as a result of activation induced by thrombin and Ca2⫹
ionophore, initiated the intracellular activation of cFXIII without the involvement of proteolysis (261, 262). Similarly, elevation of intracellular Ca2⫹ and stimulation of monocytes by
angiotensin II resulted in nonproteolytic activation of FXIII,
which catalyzed the dimerization of angiotensin receptor 1
(AT1) (1) (see more details in sect. VIIB).
B. Enzymatic Reaction Catalyzed by
Activated Factor XIII: Substrate
Specificity
TGs, including FXIIIa, catalyze an acyl transfer reaction
that consists of two major steps (124, 148, 263). First, the
peptide-bound glutamine substrate forms a binary complex
with the enzyme (thioester linkage between the carboxamide group of the glutamine and the active site cysteine). The
reaction proceeds through an oxyanion intermediate, and
ammonia is released as the acyl-enzyme intermediate is
formed. In the second step, if a substrate primary amine
group is present, the acyl group is transferred to the acyl
acceptor amine through a second oxyanion intermediate.
The amine becomes attached to the ␥-glutamyl residue via
peptide (“isopeptide”) bond, and the active-site cysteine becomes deacylated. In the absence of a substrate amine, the end
result of the reaction is the deamidation of the substrate glutamine residue. If an ⑀-amino group of a peptide-bound lysine
residue is the acyl acceptor primary amine for the reaction,
⑀(␥-glutamyl)lysyl is formed and the two peptide bonds become cross-linked. A more detailed description of the reactions catalyzed by TG is given in References 147 and 148.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
943
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
and X-ray structural studies could not be carried out. The
active-site cysteine, originally buried within the catalytic
core domain, becomes unmasked during activation and
available for reaction with its substrates and also with alkylating agents (301). Hydrogen-deuterium exchange and
chemical modification of amino acid side-chains (cysteine
alkylation and lysine acetylation) provided further insight
into structural changes occurring during the activation process. Hydrogen-deuterium exchange evaluated by MALDITOF mass spectrometry showed that FXIII activation exposes the potential substrate recognition region in the catalytic core domain (residues 220 –230), while decreased
deuteration occurred in regions 98 –104 of the ␤-sandwich
and 526 –546 of ␤-barrel 1 (364). Results from chemical
modifications demonstrated that Lys156 and Lys221, the
active sites Cys314 and Cys409 (both cysteines in the catalytic core domain), as well as Cys695 in the ␤-barrel 2
become exposed upon activation by thrombin and Ca2⫹
(365). In a further study, the exposure of Cys238, Cys327,
and Lys68 was also revealed (322). Experiments using an
inhibitory analog of FXIIIa substrate K9 DON (a ␤-caseinderived nonapeptide in which the substrate Gln residue was
replaced by the Gln isostere, 6-diazo-5-oxo-norleucine;
DON) provided further insight into the changes of FXIII-A2
structure upon activation and upon the binding of the substrate analog. K9 DON inhibited FXIIIa and provided even
greater protection of the ␤-sandwich and the ␤-barrel 1
from deuteration. At the same time, the K9 DON was able
to block the alkylation of Cys409 in the catalytic core near
the dimer interface (322).
MUSZBEK ET AL.
The substrate specificity of FXIIIa is more restricted than
that of other TGs. The primary physiological substrates of
FXIIIa are fibrin and ␣2-PI (225, 324). Twenty-three additional FXIIIa substrate proteins are enumerated in the
TRANSDAB database (http://genomics.dote.hu/wiki/).
The substrate proteins can be divided into the following
categories: coagulation factors, components of the fibrinolytic system, adhesive and extracellular matrix proteins, intracellular cytoskeletal proteins, and others (TABLE 2). The interaction of FXIIIa only with part of these
proteins has physiopathological implications; these interactions will be discussed in the respective sections of this
review.
Most recently, an interesting study demonstrated that, in
addition to the TG activity of FXIIIa, FXIII, just like TG-2,
also possesses protein disulfide isomerase (PDI) activity
944
(194). Although both activities are located on FXIII-A, they
are distinct. FXIII without being transformed into an active
TG as well as active-site alkylated FXIIIa exert PDI activity.
The physiological importance of PDI activity of FXIII still
needs to be explored.
C. Inactivation of Activated Factor XIII
Within the Clot
Extensive cross-linking of fibrin increases the resistance of
fibrin clot to fibrinolysis (103, 104). This finding suggests
that noncontrolled cross-linking by FXIIIa results in overcross-linked fibrin, and extensive cross-linking of other
plasma proteins to fibrin, that could lead to undesired prolonged persistence of thrombi. A gradual decrease of FXIIIa
activity, which has been demonstrated within the fibrin clot
and in an experimental pulmonary emboli model (catalytic
half-life: ⬃20 min) (315), suggests the existence of inactivation mechanism(s) that operate in whole plasma clot and
in thrombi. The inactivation of FXIIIa is much less investigated than its activation. Activated clotting factors are inactivated by two mechanisms (70). Proteolytically active
factors are inhibited by specific serine protease inhibitors,
serpins (like antithrombin III or tissue factor pathway inhibitor), or by less specific protease inhibitors, like ␣2-macroglobulin. The other way of inactivation of active factors is
performed by proteolytic enzymes. The latter can be carried
out by highly specific proteases, like the cleavage of activated factor V (FVa) and factor VIII (FVIIIa), by activated
protein C, or by a proteases with much broader substrate
specificity, like plasmin. Plasmin degrades fibrin, fibrinogen, FVa, and FVIIIa. As no plasma protein inhibitor of
FXIIIa has been discovered, one has to consider the proteolytic inactivation mechanism. The fibrinolytic enzyme plasmin could be a possible candidate for such a role. However,
both plasma and cellular FXIII and their activated forms are
highly resistant to plasmin (310). In addition, the powerful
inhibitory effect of fibrin-linked ␣2-PI also makes it unlikely
that plasmin could effectively be involved in the inactivation of FXIIIa in the fibrin clot. In contrast, polymorphonuclear granulocytes (PMNs) became activated and released proteases [for instance, elastase, cathepsin G, and
matrix metalloproteinase (MMP-9)] in the fibrin clot. The
released PMN proteases proteolytically degraded both
FXIII subunits and inactivated FXIIIa within the fibrin clot
(29). Such a mechanism also operated in clots made from
whole plasma, and the main physiological inhibitor of
PMN proteases, ␣1-antitrypsin, provided only limited protection. The time course of FXIIIa degradation by PMNs
suggests that the proteolytic degradation of FXIIIa by the
released proteases does not interfere with the initial crosslinking events; however, it could prevent the formation of
over-cross-linked plasma-clot and facilitate the elimination
of fibrin when it is no longer needed. It would be interesting
to investigate if this mechanism also plays a role in other
processes, in which FXIII is implicated and PMN leukocytes
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
As glutamine-containing peptides with different primary
structure around the reactive Gln residue could serve as acyl
donor substrate for FXIIIa-catalyzed TG reaction (80, 296,
356) (http://genomics.dote.hu/wiki/), the substrate-enzyme
interaction could only be described for individual peptides.
Oligopeptides corresponding to the NH2-terminal amino
acid sequence of Asn1-␣2-plasmin inhibitor (Asn1-␣2-PI),
an isoform of ␣2-PI, the major plasmin inhibitor and an
excellent glutamine substrate of FXIIIa (see sect. IV for
details), are the most well-studied acyl donor peptide substrates in this respect. Enzyme kinetic and NMR experiments with such 15-mer and 12-mer oligopeptides and with
their chemical mutants shed some light on the role of amino
acid side chains around the Gln2 substrate site in the substrate-enzyme interaction (63, 64, 232, 296). Kinetic experiments with peptides, in which Asn1 residue was either
truncated or replaced by Ala, and proton NMR analysis of
FXIII-A2*-N1-␣2-PI(1–12) complex demonstrated that
Asn1 residue is essential for effective enzyme-substrate interaction (296). Further studies demonstrated the supportive role of Glu3, Gln4, and Lys12 side chains in the reaction
(64, 296). It was shown by solution NMR methods that the
reactive peptide glutamines are encountering a distinctive
environment within the FXIIIa active site and the glutamines and residues located COOH-terminally come in
direct contact with the enzyme and adopt an extended conformation (232, 296). It is interesting that the replacement
of Gln4 by Ser or Leu doubles the catalytic efficiency of the
reaction, indicating that the FXIIIa subsite for this residue
must be relatively broad (63). Experiments with COOHterminally truncated peptides proved that amino acid residues 7–12 are essential for the interaction of N1-␣2-PI(1–
12) with the enzyme and suggested the existence of a secondary binding site on FXIII-A2*. Hydrophobic residues,
particularly Leu10, and the COOH-terminal Lys12 seemed
especially important in this respect, and direct interaction
between hydrophobic COOH-terminal residues and FXIIIA2* has been demonstrated by STD NMR (296).
MULTIPLE FUNCTIONS OF FACTOR XIII
Table 2
Protein
Coagulation factors
Substrate proteins of activated factor XIII
Substrate
Sequence Around Reactive
Gln
Fibrin(ogen) ␣ chain
77, 234, 312, 348
2, 60
105, 140, 372
181, 204
49, 212
37, 38
367
229
328, 347
99, 238
168, 349
27
65
69
1
298
39
249
ND, not determined.
accumulated. For instance, wound healing could be a target
of such an investigation.
IV. THE ACTION OF ACTIVATED FACTOR
XIII ON FIBRIN: ITS ROLE IN THE
REGULATION OF FIBRINOLYTIC
PROCESS
FXIIIa cross-links fibrin ␥- and ␣-chains into ␥-chain dimers
and ␣-chain polymers, respectively. (For detailed information on fibrinogen and fibrin clot structure, see References
93, 250, and 376.) The extremely rapid ␥-dimer formation
is the result of reciprocal intermolecular bond formation
between ␥406 lysine of one ␥-chain and ␥398/399 glutamine residue of another aligning ␥-chain (60). Cross-linking of ␣-chains, a much slower process, occurs among multiple glutamine and lysine residues, resulting in ␣-oligomers
and high Mr ␣-polymers. A small amount of ␣-␥-chain heterodimers and ␥-chain trimers or tetramers have also been
identified in fibrin (251, 343), and their presence was related to increased resistance to fibrinolysis (344). Although
␣-chain cross-linking confers the final stability to the fibrin
clot allowing strength, rigidity, and resistance to fibrinolysis
(103, 106), ␥-chain dimerization also contributes to clot
stiffness (117, 353). It has been reported that ␥A␥= fibrin
becomes more extensively cross-linked by activated FXIII
than fibrin containing only ␥A homodimers (97, 247, 248).
In sharp contrast to these findings, Siebenlist et al. (345)
found that the overall rate of ␣-polymer formation was
slower for ␥A␥= fibrin than for ␥A␥A fibrin and suggested
that the presence of ␥=-chain suppresses FXIII-induced fibrin cross-linking.
An interesting question is whether cFXIII, present in platelets incorporated into thrombi, contributes to the crosslinking of the fibrin clot. Elevated platelet count and pFXIII
concentration both increase the formation of highly crosslinked ␣-polymers (104). In the presence of normal platelets, the fibrin clot formed from FXIII-free fibrinogen becomes cross-linked (308); however, the contribution of
platelet FXIII, compared with the contribution of plasma
FXIII, was only negligible (132). Platelet FXIII is neither
secreted nor exposed to the platelet surface during platelet
activation (167, 265). The tiny amount of pFXIII, taken up
by platelets and packed into the ␣-granules, could be released during clot formation (233), but its involvement in
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
945
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
-S-Q221-L-Q-K-V-P-P-Q237-M-R-M-E-L-N-Q328-N-P-G-S-P-G-Q366-W-H-S-E-S-G-Q398-Q399-H-H-L-G-GFibrin(ogen) ␥ chain
Factor V
N.D.
␣
-Plasmin
inhibitor
(
␣
-antiplasmin)
-N-Q
-E-Q-V-S-PFibrinolytic proteins
2
2
2
Lipoprotein A
ND
Plasminogen
-S-Q322-V-R-W-E-Y
Procarboxypeptidase B/U
-F-Q2-S-G-Q5-V-L-A-A-L-S-Q292-H-I-V-F-P(thrombin-activable fibrinolysis
inhibitor; TAFI)
Adhesive/matrix proteins Thrombospondin
ND
Vitronectin
-E-Q93-T-P-V-L-KFibronectin
-A-Q3-Q-I-V-Q-POsteopontin
-S-Q34-K-Q36-N-L-L-A-PCytoskeletal proteins
Vinculin
ND
Actin
-H-Q41-G-V-M-V-GMyosin
ND
-L-Q315-L-L-K-Y-IOthers
AT1 receptor
Semenogelins I and II
ND
Protein synthesis initiation factor
ND
5A
␣2-Macroglobulin
-Q-Q670-Y-E-M-H-G-
Reference Nos.
MUSZBEK ET AL.
the cross-linking reaction is insignificant. Platelets, however, have another effect on the cross-linking reaction that is
independent of their FXIII content. When fibrin cross-linking was induced by highly purified plasma FXIII, the formation of ␣-chain polymers and the incorporation of ␣2-PIfibrin ␣-chain heterodimer into these polymers was significantly accelerated by intact FXIII-free platelets obtained
from patients with severe FXIII-A deficiency (132). The results indicate that, in physiological conditions, platelets promote the cross-linking reaction by providing a catalytic surface, on which the cross-linking of fibrin polymers is enhanced. In addition, activated platelets bind pFXIII through
fibrinogen (265) and could deliver it to the thrombus.
Cross-linking of fibrin chains by FXIIIa, in the absence of
␣2-PI, decreased the rate of lysis induced by exogenous plasmin (103). FXIII, in excess to its normal plasma concentration, increased the extent of fibrin cross-linking and inhibited plasmin-induced degradation of fibrin (103, 104).
However, the lysis rate for cross-linked fibrin prepared
from plasma was much slower than the lysis rate of crosslinked fibrin prepared from purified fibrinogen, and 10-fold
higher plasmin concentration was needed to bring the lysis
rate of plasma-derived fibrin to that of fibrin formed from
purified fibrinogen. This finding emphasizes that plasma
component(s) incorporated into the clot are required for the
effective protection of fibrin from fibrinolysis. The impaired
binding of plasminogen to cross-linked fibrin (237, 326) might
also play a role in the FXIII-induced resistance to fibrinolysis.
␣2-PI, the main physiological inhibitor of plasmin, is an
excellent substrate of FXIIIa, and it becomes cross-linked to
fibrin ␣-chains during the initial phase of cross-linking reaction (324). It is synthesized in the liver and becomes secreted as a protein of 491 amino acids starting with a methionine (Met1-␣2-PI). In the plasma, a protease, antiplasmin cleaving enzyme (APCE), cleaves off the NH2-terminal
propeptide of 12 amino acid residues (202) and transforms
Met1-␣2-PI into Asn1-␣2-PI (32, 357). The transformation
is only partial; in plasma, the ratio of Met1-␣2-PI to Asn1␣2-PI is 3:7. Only the Asn1-␣2-PI isoform is a good substrate for FXIIIa (357). It provides a Gln residue penultimate to Asn1 (Gln2; Gln14 in Met1-␣2-PI) to the crosslinking reaction. The same Gln residue is sheltered by the
NH2-terminal 12-residue peptide of Met1-␣2-PI and makes
it inaccessible for FXIIIa (203). In fact, the addition of only
three COOH-terminal amino acids of the propeptide makes
Asn-␣2-PI a poor substrate for the TG reaction (136). With
the use of recombinant Gln2Ala ␣2-PI mutant, three further
946
During the formation of the fibrin clot, the Gln2 site becomes cross-linked to Lys303 residue of ␣-chain of fibrin
(181). It is interesting that this Lys residue is not involved in
the formation of ␣-chain polymers. The formation of ␣2-PIfibrin ␣-chain heterodimers is a rapid reaction; it only
slightly lags behind fibrin ␥-chain dimerization and precedes fibrin ␣-chain cross-linking. It can also be cross-linked
to fibrinogen, although at a significantly slower rate. The
existence of ␣2-PI cross-linked to plasma fibrinogen has also
been reported (252). There is no acyl acceptor lysyl residue
in the ␣2-PI molecule; therefore, no ␣2-PI dimer is formed
(360). Only Asn1-␣2-PI is cross-linked to fibrin, which explains that part of ␣2-PI remains in the serum. Cross-linked
␣2-PI retains its full inhibitory activity and makes fibrin
strikingly resistant to digestion by plasmin (202, 325). The
concerted action of FXIII and ␣2-PI plays a key role in the
protection of fibrin clots against prompt elimination by fibrinolysis. The dodecapeptide with the NH2-terminal sequence of Asn1-␣2-PI was shown to inhibit the cross-linking of ␣2-PI to fibrin, resulting in accelerated fibrinolysis
(182). ␣2-PI deficiency is a very rare, but severe bleeding
diathesis with symptoms similar to those of FXIII-A deficiency (56). Phenylglyoxal inactivated form of ␣2-PI (205)
or the active site Arg364Ala mutant (206) enhanced urokinase-induced plasma and whole blood clot lysis significantly by competing with the native ␣2-PI for FXIIIa catalyzed incorporation into fibrin. Jansen et al. (158) reported
that the tissue plasminogen activator (tPA)-induced lysis of
whole blood clot from a patient with ␣2-PI deficiency was
not or only slightly increased by an antibody that inhibited
FXIIIa. This finding suggests a primary role of ␣2-PI to
fibrin cross-linking in the reduction of clot lysis by FXIII.
However, Sakata et al. (326) demonstrated in ␣2-PI depleted plasma, and in purified system without ␣2-PI, that
fibrin cross-linking itself and the consequent decrease in the
binding of plasminogen to fibrin also contribute to the reduction of fibrinolysis by FXIII.
The above findings clearly prove that FXIII has a major role
in protecting fibrin from degradation by the fibrinolytic
system. Most recently, the protective effect of FXIII against
fibrinolysis was also demonstrated in clots formed under
flow (264). Fibrin cross-linking and the binding of ␣2-PI to
fibrin by FXIIIa might work together or consecutively. Using a unique antibody, which inhibited FXIIIa-fibrin and
FXIIIa-fibronectin interaction but did not influence the
cross-linking of ␣2-PI to fibrin, McDonagh and Fukue (235)
demonstrated that both ␣2-PI-fibrin cross-linking and
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
FXIII may influence fibrinolysis by two mechanisms: 1) the
cross-linking of fibrin ␣-chains into high-Mr ␣-polymers
may render the clot more resistant to fibrinolysis, and 2) the
binding of ␣2-PI, and perhaps also other plasma components, to fibrin may protect the fibrin clot and prevent its
prompt elimination by the powerful fibrinolytic system.
glutamine donor sites were found in the molecule at positions 21, 419, and 447. However, even in the mutant protein,
their kinetic efficiency was much lower than that of the Gln2
residue (204). It has been shown that the Arg6Trp polymorphism in Met1-␣2-PI affects its cleavage rate by APCE in favor
of the Arg6 variant, but this polymorphism does not affect the
poor cross-linking of Met1-␣2-PI to fibrin (203).
MULTIPLE FUNCTIONS OF FACTOR XIII
␣-chain polymerization were necessary to regulate the rate
of fibrinolysis. In experimental pulmonary embolism, both
FXIIIa-mediated fibrin cross-linking and ␣2-PI-fibrin crosslinking increased the resistance to endogenous and tPAinduced fibrinolysis (307). In summary, the above studies
indicate that the cross-linking of ␣2-PI to fibrin occurs at an
early stage of thrombus formation and provides protection
of the newly formed fibrin clot against the prompt elimination
by the fibrinolytic system, while the increased resistance to
thrombolysis of matured thrombi could be the consequence of
the much slower, extensive cross-linking of ␣-chains, including ␣-chains to which ␣2-PI had been attached.
Thrombin activatable fibrinolysis inhibitor (TAFI; also
known as procarboxypeptidase U, procarboxypeptidase B,
or procarboxypeptidase R) plays an important role in the
regulation of fibrinolysis (71, 253, 311). TAFI is activated
by thrombin, a process that is greatly enhanced in the presence of thrombomodulin. Activated TAFI inhibits fibrinolysis by abrogating the following plasmin-mediated positive-feedback mechanisms of fibrinolysis. Once formed,
plasmin starts to digest the clot by catalyzing cleavages after
selected arginine and lysine residues. These cleavages provide extra binding sites to Glu plasminogen and tPA by
exposing COOH-terminal lysine residues. This way the cofactor activity of fibrin in Glu plasminogen activation increases. In addition, the partially cleaved fibrin serves as a
cofactor for the conversion of Glu plasminogen into Lys
plasminogen by plasmin. The activation of Lys plasminogen by tPA is ⬃20-fold more efficient than the activation of
Glu plasminogen. TAFI interferes with these positivefeedback mechanisms by removing the newly exposed
COOH-terminal lysine residues of degraded fibrin, and
thereby diminishing the catalytic efficiency of plasmin
formation. TAFI has been shown to be a substrate for
FXIIIa (367), which catalyzed the polymerization of
TAFI and its cross-linking to fibrin. TAFI contains both
acyl donor Gln and acyl acceptor Lys residues; Gln2,
Gln5, and Gln294 are the preferred acyl donor sites. It
has been suggested that the cross-linking may facilitate
the activation of TAFI, stabilize the enzymatic activity,
FXIIIa catalyzed the incorporation of labeled amines and a
Gln-containing substrate peptide into plasminogen, demonstrating that it contains both glutamine and lysine donor
sites (37). FXIIIa can polymerize plasminogen into high Mr
multimers and form fibronectin-plasminogen heteropolymers. The physiological significance of these findings, if
any, remains to be elucidated. The protein part of lipoprotein (a) [Lp(a)], an atherogenic risk factor, strikingly similar
to plasminogen, also contains acyl donor glutamine residues, for the FXIIIa catalyzed TG reaction (49). In native
Lp(a), part of the substrate glutamyl sites are buried and
inaccessible for FXIIIa. Lp(a) competes with plasminogen
and tPA for binding sites on fibrin(ogen) and inhibits plasmin generation (227). FXIIIa cross-links Lp(a) to fibrinogen
(317), but the concentration of FXIII in this experiment was
well above normal plasma concentration, and the incubation times were too long to support the physiological relevance of this interaction.
In summary, FXIII is an essential component of the clotting
system; its deficiency causes severe bleeding diathesis. The
main function of FXIII in hemostasis is the stabilization of
newly formed fibrin. By cross-linking fibrin ␥-chains into
dimers and fibrin ␣-chains into high-molecular-weight
polymers, FXIII increases the rigidity and strength of the
fibrin clot and protects it against shear stress in the circulation. FXIII also protects fibrin from the prompt elimination
by the fibrinolytic system. This mechanism involves the covalent cross-linking of ␣2-PI, the major plasmin inhibitor,
to fibrin strands, the diminished binding of plasminogen to
cross-linked fibrin, and the reduced lysis of cross-linked
fibrin by plasmin.
V. FACTOR XIII IN WOUND HEALING AND
ANGIOGENESIS
A. Clinical Data and Animal Experiments
Supporting the Involvement of Factor XIII
in Wound Healing
The first FXIII-deficient patient, recognized by Duckert et
al. in 1960, demonstrated impaired wound healing and abnormal scar formation, in addition to the severe bleeding
complications (94, 331). The frequency of poor wound
healing in severe FXIII deficiency was reported between
14 –36% in various reviews (22, 44, 154, 336). The essential role of FXIII in the wound healing process was clearly
proven in transgenic mice, in which the targeted deletion of
FXIII-A gene caused complete FXIII-A deficiency (199). In
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
947
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
There are also other interactions between FXIII and components of the fibrinolytic system. It has been shown that
plasminogen activator inhibitor 2 (PAI-2), an inhibitor of
urokinase-type plasminogen activator (uPA), is cross-linked
to fibrinogen or fibrin by FXIIIa (163, 313). Cross-linked
PAI-2 retained its activity and effectively inhibited uPAmediated fibrin clot lysis. PAI-2 is detectable in the plasma
only during pregnancy; although it is mainly of cellular
localization, peripheral blood monocytes incorporated into
the fibrin clot are capable of secreting it. The cross-linking
of PAI-2 to fibrin may serve compartmentalization of the
inhibitor. Monocytes incorporated into thrombus could
therefore play a role in the downregulation of plasminogen
activation (313).
and protect the active enzyme from further degradation.
However, no experimental support has been provided to
this hypothesis, and the long incubation time required for
the cross-linking questions the physiological implications
of this mechanism.
MUSZBEK ET AL.
the FXIII-A-deficient mice, the healing of excisional wound
was considerably delayed (FIGURE 8); incomplete reepithelization, persisting necrotized fissure, and abnormal scar
formation were observed at day 11 after excision, when
wild-type mice demonstrated complete closure of the
wound (150). Substitution with human pFXIII restored the
normal healing process (FIGURE 8).
The influence of FXIII polymorphisms on the prevalence
and presentation of CVLU has been intensively studied by
Gemmati and co-workers (111-113, 361) FXIII-A Val34Leu,
Tyr204Phe, Pro564Leu, and FXIII-B Arg95His did not influence the prevalence of CVLU (361). However, the presence of Leu34 and Leu564 alleles was associated with
smaller ulcer surface, and the size of ulcer inversely related to the number of Leu34 alleles (111, 113, 361).
Healing time was increased in patients with the Val34Val
genotype compared with Leu34 carriers (111). Furthermore, FXIII-A Leu34 beneficially influenced the healing
time after superficial venous surgery (112). The above
results suggest that FXIII is involved in the pathomechanism of the healing process of venous leg ulcer, its topical
application is beneficial, and the accelerated rate of FXIII
activation and altered fibrin structure in FXIII-A
Val34Leu polymorphism decrease the severity of the disease. Further details on this topic can be found in an
excellent recent review (386).
Early sporadic reports demonstrated the beneficial effect
of FXIII supplementation on the healing of surgical
wounds in postoperative situations (114, 115, 246).
In inflammatory bowel disease, ulcerative colitis, and
Crohn’s disease, the compensation of consumption and
loss of FXIII into inflamed tissue by substitution seems to
improve the healing tendency (133, 226, 337). Treatment
948
When acquired FXIII deficiency was induced in rats by CCl4
treatment, the healing of microvascular anastomosis and
burn injury were impaired, but the healing process could be
normalized by the administration of FXIII (153, 284). The
healing of osteotomies in sheep could also be stimulated by
FXIII, and the bone formed in this condition had higher
tensile strength (62).
B. Effect of Factor XIII on Fibroblasts,
Macrophages, and Extracellular Matrix
Fibrin gel formation, macrophage invasion, fibroblast migration/proliferation, production of extracellular matrix,
and angiogenesis are the key events in wound healing. Earlier and recent studies enlightened certain details of the
involvement of FXIII in these mechanisms. FXIII exerts its
effect by cross-linking extracellular matrix proteins, by affecting fibroblasts and macrophages, and by promoting the
complex mechanism of angiogenesis.
The first publication in this field came directly after discovering the first FXIII-deficient patient. Beck et al. (36) demonstrated that growth of fibroblasts in the plasma of this
patient was impaired, the contact between cells was lost,
their normal elongated cell shape became irregular, and the
cells failed to produce collagen fibers. Supplementation
with normal plasma or with partially purified FXIII corrected the defects; however, a higher amount of FXIII was
required for the correction of growth impairment than for
restoring normal fibrin cross-linking. The enhancement of
fibroblast proliferation by cross-linking fibrin matrix (366)
and by direct proliferation-inducing effect of FXIII (54)
were both reported in early preliminary studies. Another
study indicated that FXIII binds to the extracellular matrix
and matrix assembly sites, where it remains active (34). Its
binding to a putative cell surface receptor, which mediates
internalization and degradation, was also suggested. The
migration of fibroblasts into fibrin gel was greatly enhanced
by extensive cross-linking of fibrin ␣-chains, while the
dimerization of ␥-chains was not sufficient to support fibroblast migration (53). Fibronectin, an important component
of extracellular matrix, is cross-linked by FXIIIa at cellular
matrix assembly sites (35), which complements disulfidebonded multimer formation in the stabilization of assembling fibronectin molecules. Fibronectin covalently incorporated into the fibrin clot by FXIIIa supports fibroblast
adherence and promotes the migration of cells into the gel
(126). Cell attachment and spreading on composite fibrin/fibronectin clot significantly decreased when recombinant fibronectin with mutations in the FXIIIa substrate
glutamine sites was used, i.e., fibronectin must be crosslinked to fibrin by FXIIIa to ensure maximal cell attachment (74). In addition to fibrin and fibronectin, several
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
Chronic venous leg ulcer (CVLU) represents a peculiar form
of impaired wound healing with multicausative background. The possible involvement of FXIII in the pathomechanism of this disease is of particular interest. An earlier
study demonstrated decreased FXIII level in a small group
of patients with CVLU (369), while in a more recent report
FXIII-A antigen level did not differ significantly between
CVLU patients and matched controls (n ⫽ 91 in both
groups) (113). However, in the latter study, 11% of the
cases had moderate FXIII deficiency (FXIII-A antigen
ⱕ0.65 U/mL). Although only a relatively small number of
patients were recruited for the four studies evaluating the
effect of topical treatment of CVLU with FXIII concentrate,
the studies unanimously reported a beneficial effect (131,
134, 297, 380). Accelerated wound surface reduction,
shorter healing time, improved availability of granulation
tissue, and decreased secretion and bleeding tendency were
observed following daily local administration of the concentrate in CVLU, but not in arterial-venous mixed disease
(380).
with recombinant FXIII also improved established experimental colitis in rats (83).
MULTIPLE FUNCTIONS OF FACTOR XIII
other adhesive proteins, vitronectin, osteopontin, thrombospondin, and von Willebrand factor, are substrates of
FXIIIa (TABLE 2). Their cross-linking to other connective
tissue matrix components might have further impact on
the attachment of cells to the matrix and on the healing of
injuries.
Macrophages have a well-known important role in the
wound repair process (3, 118, 119). FXIII of tissue macrophages might play a role in maintaining and modulating the
connective tissue matrix in normal conditions. This hypothesis, however, still lacks experimental support. In theory,
FXIII of the extracellular compartment might act on monocytes/macrophages by influencing their activation, differentiation, and migration, while cFXIII present in these cells
might play a role in the mechanism of macrophage migration and phagocytosis. Like in the case of fibroblasts, FXIIIa
enhanced the proliferation of peripheral blood monocytes,
accelerated their migration through the filter of a Boyden
chamber by twofold, and significantly inhibited monocyte
apoptosis (85). These changes were related to the downregulation of thrombospondin-1 and to the upregulation of
c-Jun and Egr-1 in monocytes. Active-site blocked FXIIIa
was ineffective, suggesting that the TG activity of FXIIIa is
essential for its effects on monocytes (85). The above results
seem to be in contradiction to the finding that anti-FXIII-A
antibodies did not influence the binding of THP-1 cells (a
human monocytic leukemia cell line) to fibronectin and
their migration on fibronectin surface (16). However, the
leukemic cells might behave differently from normal monocytes, and the inhibitory effect of the antibody was not
proven in this publication. It was also shown that in fibrin
gel containing cross-linked fibronectin the migration of
monocytes slowed down (197). Although the relation of
migration in fibrin gel to the migration required for wound
healing is questionable, the adhesion to cross-linked fibronectin might represent an anchoring mechanism that
keeps macrophages where they are needed for the wound
healing process.
The involvement of monocyte/macrophage cFXIII in the
wound healing process has not been explored to a full
extent. FXIII-A2 cannot be secreted by the classical pathway, and there is no proof for its secretion by other
secretory mechanisms. Nevertheless, leakage of cFXIII from
damaged cells might contribute to the local cross-linking of
proteins. The activation of monocytes/macrophages by the
alternative pathway, for instance, by interleukin-4, highly increased FXIII-A expression (362). Alternatively activated or
M2 macrophages express scavenger receptors, bind apoptotic
cells, and are important in mediating wound healing following
tissue damage (96). cFXIII is involved in receptor-mediated
phagocytosis by monocytes (6, 329). Removal of cell debris
and apoptotic cells is an important step in the wound healing process, and the decreased phagocytosis of monocytes/
macrophages in FXIII-A deficient patients might contribute
to the impaired tissue repair. Another intracellular mechanism involving cFXIII in wound healing could operate
through the angiotensin II-induced release of monocytes
from their splenic reservoir and migration to the site of
injury (358). Binding of angiotensin II to its receptor (AT1)
could lead to cFXIII activation and consequently to the
covalent dimerization of AT1 by FXIIIa, which induces the
activation and mobilization of monocytes (1). (The expression and activation of cFXIII in activated monocytes and
the involvement of cFXIII in the function of these cells are
detailed in section VIIB.)
C. Proangiogenic Effect of Activated Factor
XIII
Angiogenesis is a complex process that includes remodeling and sprouting of new capillaries from existing
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
949
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
A recent study confirmed and extended sporadic preliminary results on the direct effect of FXIII on fibroblasts using
more elaborate techniques; this study also shed light on the
mechanism by which FXIII exerts its effect (85). FXIIIa, but
not nonactivated FXIII or inactivated FXIIIa, bound to skin
fibroblast, enhanced their migration in a wound-migration
assay, increased the incorporation of [3H]thymidine into
the cells, and decreased fibroblast apoptosis. The finding
that an antibody against ␣v␤3-integrin, the vitronectin receptor, blocked the binding of FXIIIa to fibroblasts implicated this receptor in the effect exerted by FXIIIa (85). In
addition, FXIII counteracted the toxic effect of metalloproteinases on dermal fibroblasts, although it is not clear from
this study if FXIII added to the cell culture became activated
or not (385).
Another interesting aspect of the relationship between
macrophages and FXIII is the FXIII-dependent generation of a complement C5-derived monocyte chemotactic
factor during clotting (286). The major monocyte chemotactic factor identified in the serum had a Mr (75 kDa)
by gel filtration that is much higher than that of the active
complement fragment C5a or C5a des-Arg (8 kDa). This
finding suggests that C5a was cross-linked to another
plasma protein, although no direct evidence has been
provided. Most recently, a plasma protein with the features of ribosomal protein S19 (RP S19) was identified,
which was dimerized by FXIIIa on the surface of activated platelets during coagulation and converted into a
monocyte-selective chemoattractive factor (339). Interestingly, RP S19 showed immunological cross-reactivity
with C5a. However, the Mr of the active dimeric form
was only 32 kDa. It remains to be seen what is the exact
relationship between the cross-linked C5a fragment and
the active ribosomal protein S19-like factor. Such monocyte specific chemotactic factor(s) formed during clotting
could be involved in recruiting monocytes/inflammatory
macrophages to the site of injury.
MUSZBEK ET AL.
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
950
MULTIPLE FUNCTIONS OF FACTOR XIII
blood vessels. It is involved in a wide range of physiopathological processes from wound healing to tumor
growth. The results published in the last decade strongly
suggest that the role of FXIII in tissue repair and wound
healing is at least in part mediated by its proangiogenic
effect. The interaction of FXIII with immortalized human
endothelial cells was first reported a decade ago by Dallabrida et al. (84) and by Dardik et al. (89). The binding
itself did not require TG activity, nonactivated FXIII,
FXIIIa, and inactivated FXIIIa all bound to the cells via
␣v␤3 integrin (vitronectin receptor) and activation of the
receptor by MnCl2 highly increased the binding (84, 89).
In addition, FXIII seems to be capable of mediating endothelial cell-platelet interaction (89).
Dardik et al. (90) also demonstrated in the Matrigel tube
formation model that FXIIIa caused a dose-dependent enhancement array formation of HUVECs. No effect was observed on the addition of FXIII or iodoacemide inhibited
FXIIIa, or when the TG activity of FXIIIa was blocked by a
polyclonal antibody (90). Interestingly, an opposite effect
was observed by Dallabrida et al. (84) using a simian virus
40 (SV40) transformed human dermal microvascular endothelial cell line (HMEC-1) and fibrin gel. Differences in the
features of the two cell types, e.g., in their TSP-1 content
The role of FXIII in angiogenesis has also been verified in
several in vivo animal models. Injection of FXIIIa into the
cornea subepithelially induced the formation of a rich network of capillaries within 36 h. At the same time, the positive staining for TSP-1 observed in the fibers of the stroma
and in keratocytes of nontreated cornea disappeared (90).
Topical application of FXIIIa produced a significantly
higher number of new vessels in neonatal heterotopic
mouse heart allograft model and increased the contractile
performance of the allograft compared with untreated normal mice (86). Here again, FXIIIa decreased TSP-1 mRNA
and protein expression. These findings also demonstrate
that FXIIIa exerts its proangiogenic effect, at least in part,
through the downregulation of TSP-1. In FXIII-A knockout
mice, the formation of new vessels in subcutaneously injected
Matrigel plug dramatically decreased compared with normal
mice (86). If FXIIIa was added to the Matrigel, the number of
vessels almost reached the normal level. When a critical-sized
bone defect in the tibia of rat was filled with hydroxyapatite,
the addition of FXIII stimulated the in-growth of microvessels
into the implant (180). Although FXIII added to hydroxyapatite was not preactivated, it was transformed somehow into
its active form locally as demonstrated by staining with an
antibody specific for the active form.
The mechanism of the proangiogenic effect of FXIIIa was
partly enlightened in Inbal’s laboratory. FXIIIa bound to
␣v␤3 receptor enhances its noncovalent interaction with
VEGFR-2, and even resulted in partial cross-linking between the ␤3 subunit of vitronectin receptor and
VEGFR-2 (88). FXIIIa then induced tyrosine phosphorylation and activation of VEGFR-2 that led to the phosphorylation and activation of Akt and the extracellular
signal-regulated protein kinases (ERKs) and to the upregulation of the transcription factors c-Jun and Egr-1.
Akt promotes growth factor-mediated cell survival and
has been implicated in angiogenesis. The activation of
ERKs together with the zinc-finger transcription factor,
Egr-1-mediated pathway promotes cell proliferation. As
mentioned above, the downregulation of TSP-1 is a key
event in the proangiogenic effect of FXIIIa. Overexpres-
FIGURE 8 Impaired wound healing in FXIII-A knock out mice. A: representative views of the gross appearance of excisional wounds of control,
FXIII-A-deficient, and FXIII-A-deficient/pFXIII-treated mice on days 0, 8, and 11. B: follow up of wound scores at different time points of control
(squares), FXIII-A-deficient mice (open triangles), and FXIII-A-deficient/pFXIII-treated mice (black triangles). Each group included 10 animals. Data
are presented as means ⫾ SD. Mean ⫾ SD area under the curve (AUC) was calculated from the wound scores. AUC was significantly higher
in either the control or FXIII-treated groups compared with the FXIII-A-deficient animals (P ⬍ 0.001 in both cases). [From Inbal et al. (150), with
permission.]
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
951
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
Although the binding of the nonactivated form might be
important for targeting FXIII to endothelial cells, its active form is required to exert proangiogenetic effects.
FXIIIa, but not the zymogen FXIII or active-site blocked
FXIIIa, enhanced human umbilical endothelial cell
(HUVEC) migration in two different migration assays
(90). FXIIIa also increased [3H]thymidine incorporation
into HUVECs and significantly suppressed the apoptosis
of these cells as demonstrated in TUNEL assay (90).
FXIIIa had no effect on the secretion of vascular endothelial growth factor (VEGF) into the medium, and neither VEGF protein level nor steady-state VEGF receptor-2 (VEGFR2) mRNA levels were affected by FXIIIa. In
contrast, FXIIIa induced an almost complete disappearance of thrombospondin-1 (TSP-1) mRNA from the cells
and markedly reduced the amount of TSP-1 in the conditioned medium. TSP-1, a homotrimeric multifunctional glycoprotein produced and secreted by many cells,
is involved in the inhibition of angiogenesis by inhibiting
endothelial cell migration and proliferation and by inducing apoptosis (200, 245, 276). The interference of
TSP-1 with wound healing-associated angiogenesis was
also demonstrated in transgenic mice with targeted overexpression of TSP-1 in the skin (355).
(90), opposing response in VEGFR-2 and VEGF expression
to ␣v integrin suppression (87), might be partly responsible
for the discrepancy. Another significant difference was the
composition of gel matrix; Matrigel, a gelatinous protein
mixture secreted by Engelbreth-Holm-Swarm (EHS) mouse
sarcoma cells, resembles the complex extracellular environment found in many tissues as opposed to fibrin gel consisting of a single component.
MUSZBEK ET AL.
sion of c-Jun in HUVECs and in cardiac allografts has
been linked to the downregulation of TSP-1 (86, 88), and
this effect is mediated by Wilm’s tumor-1 (WT-1), another transcription factor (88). As the level of WT-1 is
not influenced by FXIIIa, its effect seems to be exerted
through the modulation of WT-1 to TSP-1 association.
VI. OTHER EFFECTS OF FACTOR XIII
A. Role of Factor XIII in Maintaining
Pregnancy
The levels of fibrinogen, factors VII, VIII, IX, and X progressively increase during pregnancy (354). A similar pattern is followed by FXIII-B but not FXIII-A. FXIII-A concentration and consequently also FXIII-A2B2 concentration
and FXIII activity start to decrease between the fifth to
eighth gestational week (wG) and following a gradual reduction reaches ⬃50% of normal average (41, 73, 129,
277, 368). FXIII then rises with the onset of labor and falls
FIGURE 9 A hypothetic scheme for the involvement of FXIII in wound healing. pFXIII activated in the coagulation pathway plays a pleiotropic role that involves antifibrinolytic effect, modulation of extracellular matrix,
promotion of monocyte/fibroblast migration, proliferation, survival, and a complex effect on endothelial cells
leading to angiogenesis. cFXIII in monocytes/macrophages recruited to the site of injury becomes upregulated
and activated. Activated cFXIII of monocytes/macrophages promotes phagocytosis. This way it is involved in
the removal of cell debris and apoptotic cells. The evidence for individual pathways shown is discussed and
evaluated in the text.
952
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
The above findings leave no doubt that FXIIIa is a proangiogenic factor, and this function of FXIIIa is important for
the promotion of wound healing. The biochemical mechanism of FXIIIa-induced angiogenesis and its participation in
wound healing have been partly clarified (FIGURE 9), but
there are still a number of missing links that need to be
explored. Further basic and clinical research in this area
may lead to the exploitation of this proangiogenic effect in
clinical practice. Topical application of FXIII concentrate
might be used to promote the angiogenesis in poorly healing
wounds, and the inhibition of FXIIIa-induced angiogenesis
might have a potential in antitumor therapy.
MULTIPLE FUNCTIONS OF FACTOR XIII
again postpartum. Whether the reduction of plasma FXIII
during pregnancy represents decreased synthesis of
FXIII-A, increased utilization/consumption, or simple dilution by expanded plasma volume is not clear.
The concentration of pFXIII required for maintaining pregnancy has not been established. The observation that FXIIIB-deficient women, who have FXIII level in the range of
5–10%, do not experience miscarriage (116, 323) suggests
that pFXIII concentration significantly lower than normal is
sufficient to maintain pregnancy without substitution therapy. A patient with pFXIII activity, as low as 0.3%, was
able to carry two pregnancies until wG 31 and 37, and only
then did ablation of placenta with bleeding complications
occur that needed replacement therapy and caesarian sections (243). Present recommendations for FXIII substitution of FXIII-deficient patients suggest that 10% presubstitution FXIII level is ideal to prevent miscarriages (26, 139,
173). Ogasawara et al. (283) investigated FXIII level in 424
patients with a history of first trimester miscarriages. No
difference in the subsequent miscarriage rates was found
between patients with normal and moderately decreased
level of FXIII. The above findings clearly demonstrate that
pFXIII is essential to carry out normal pregnancy, but the
level required to prevent miscarriages is significantly below
the reference interval of pFXIII.
FIGURE 10 Distribution of FXIII-A in the definitive placenta. The major components, related to the role of
definitive placenta, are indicated. FXIII-A-containing cellular and structural elements are depicted in maroon.
These include macrophages in the placental villi (Hofbauer cells) and in the decidua and the Nitabuch’s fibrinoid
layer. This layer is located between the zona spongiosa and zona compacta where the release of placenta will
take place. The fibrinoid deposits contain fibrin and fibronectin cross-linked by FXIIIa.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
953
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
FXIII-A is not required for conception, but it is essential for
maintaining pregnancy; patients with severe FXIII-A deficiency and without replacement therapy experience recurrent pregnancy losses in the first trimester (see reviews in
Refs. 26, 142, 151, 173). Homozygous female FXIII-A
knock out mice were also capable of becoming pregnant;
however, most of them died of excessive vaginal bleeding
(189). Massive placental hemorrhage with subsequent necrosis developed in the uteri by day 10. As embryonic development of the FXIII-A-deficient fetus is normal, the defect should be at the maternal side. FXIII-A has been identified in histiocytes of the uterus (11), in macrophages of the
implantation tissue (25), and in the placenta (9). In the
placenta small mononuclear round-shaped FXIII-A-containing cells appear at the fifth wG and then they rapidly
proliferate and differentiate into large stellate cells reaching
⬃30% of total cell number (171). From the eighth wG they
tend to accumulate in the peripheral part of chorionic villi.
The location of FXIII-A-containing cells in the definitive
placenta is demonstrated in FIGURE 10. Although the presence of FXIII-A-containing cells both in the placenta and
the implantation tissue suggests the involvement of these
FXIII-A-containing macrophages in the development of
normal pregnancy, the simple fact that replacement by
pFXIII concentrate is sufficient to carry FXIII-A-deficient
patients successfully through pregnancy (47, 100, 185,
316) proves the primary importance of pFXIII. The function of cFXIII present in uterine and placental macrophages remains unclear; it might be involved in the crosslinking of extracellular matrix proteins in these tissues.
MUSZBEK ET AL.
It is very likely that activated FXIII cross-links and stabilizes fibrin (and fibronectin) of the Nitabuch’s layer,
and protects it from fibrinolysis. The urokinase-type
plasminogen activator (uPA) is abundant in placental
cells (98, 101) and, together with its cellular receptor and
inhibitors (PAI-1 and PAI-2), FXIII may act as a regulator of uPA-induced fibrinolysis. The fibrinoid of the Nitabuch’s layer seems to play a complex role. In addition to
its sealing effects, it also acts as a protective immunologic
“barrier” between feto-maternal tissues and participates
in anchoring the placenta. The absence of its cross-linking by FXIIIa could compromise all these functions. The
angiogenic effect of FXIIIa may contribute to the angiogenesis during placentation. Insufficient formation of cytotrophoblastic shell was also reported in woman with
FXIII deficiency (24). Although the essential role of
pFXIII in maintaining pregnancy has been firmly established, we are only beginning to understand its precise
role and the mechanisms through which it ensures normal placental development as well as prevents peri-placental bleeding, detachment of the placenta from the
uterus, and subsequent spontaneous fetal loss.
B. Effect of Factor XIII on Vascular
Permeability
The first report concerning the influence of FXIII on vascular permeability came from Hirihara et al. (135). In their
experiments the vascular permeability enhancement induced by intradermal injection of anti-endothelial cell antiserum was demonstrated by the leakage of Evans blue from
vascular compartment after 15 min and by measuring the
swelling after 1, 2, or 3 days, respectively. Local administration of pFXIII suppressed both the acute and the subacute increase of permeability significantly. It was not explored if FXIII became activated following administration
or the inactive form was sufficient to exert the inhibitory
effect. A more detailed in vitro study used porcine endothelial cell culture and ischemic-reperfused rat heart in a Lan-
954
gendorf perfusion system to address the effect of FXIII on
endothelial barrier function (275). Physiological concentration of activated pFXIII or cFXIII decreased the permeability of endothelial cell monolayer for Trypan blue-labeled
albumin in a two-compartment system separated by a porous filter. In addition, they were able to prevent the rapid
rise of macromolecule permeability that was induced by an
inhibition of endothelial mitochondrial and glycolytic energy production. Likewise, FXIIIa prevented the increase of
myocardial water content induced by low-flow ischemia
and subsequent reperfusion on rat heart (275). Nonactivated pFXIII and cFXIII, inactivated FXIIIa and FXIII-B
failed to exert any effect in these experiments, suggesting
that the effect of FXIIIa is related to its TG activity. Immunohistological investigations demonstrated the deposition
of FXIIIa at interfaces of adjacent endothelial cells and between the cells and the filter support. The mechanism of the
effect of FXIIIa on endothelial cell permeability has not
been revealed. A variety of FXIIIa substrate adhesive proteins, like fibronectin and vitronectin, reside in the intracellular clefts and subcellular matrix; their cross-linking might
influence the paracellular pathway for passage of macromolecules through the monolayer.
A few clinical and experimental animal studies also support
the effect of FXIII on vascular permeability. In a small prospective investigation on newborns and young children
with congenital heart disease undergoing open-heart surgery, peri-operative cardiac edema formation was related to
the decrease of pFXIII concentration (381). Preoperative
treatment with FXIII concentrate reduced the extent of
myocardial swelling and the extension of generalized
edema. In another small study on children with open-heart
surgery, a single intravenous dose of pFXIII reduced severe
pleural effusion (333). Czabanka et al. (82) demonstrated
that the administration of pFXIII together with endotoxin
reduced the increase of vascular permeability in the rat, and
leukocytes were not needed for this effect.
The findings published, so far, clearly indicate that FXIIIa
reduces the permeability of endothelial cell monolayers to
macromolecules. Revealing the mechanism by which FXIIIa
exerts its effect seems to be a promising and interesting field
of future investigations. A few data also suggest that FXIII
decreases the enhanced vascular permeability induced by
various stimuli in experimental and clinical setting. However, the results of these sporadic observations and their
possible clinical impact need to be confirmed by wellplanned and focused clinical and animal studies.
C. Cardiac Effect of Factor XIII
In 2006 a surprisingly novel aspect of FXIII-A deficiency
was reported by Nahrendorf et al. (268). They compared
the survival of homozygous and heterozygous FXIII-A
knock out mice (FXIII-A⫺/⫺ and FXIII-A⫹/⫺, respectively),
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
The exact role of FXIII in maintaining pregnancy and the
cause of pregnancy losses in women with FXIII deficiency is
unknown. Two substrates of FXIIIa, fibrin(ogen) and fibronectin, are important components of the Nitabuch’s fibrinoid layer, the layer in the decidua basalis between the
zona compacta and zona spongiosa, where the placenta
detaches itself from the uterus at birth (FIGURE 10). Kobayashi et al. (184) reported that FXIII-A was present in the
extracellular space of the extravillous cytotrophoblast
forming the cytotrophoblastic shell adjacent to Nitabuch’s
layer. FXIII-A colocalized with fibrinogen and fibronectin
at the Nitabuch’s layer next to the cytotrophoblastic shell
(24). The localization of FXIII-A supports its plasmatic origin, while the absence of FXIII-B at these locations suggests
that pFXIII became activated and FXIII-B dissociated from
FXIII-A.
MULTIPLE FUNCTIONS OF FACTOR XIII
were detected in the myocardium of patients with infarct
rupture (267), and moderately decreased pFXIII levels
(57 ⫾ 8%) were reported in three patients presenting with
acute myocardial rupture following myocardial infarction
(270). These interesting findings could provide a basis for
controlled clinical studies on the risk of cardiac rupture
following myocardial infarction in individuals with moderately decreased FXIII, including heterozygous FXIII-A-deficient patients.
An imbalance in extracellular matrix turnover was also observed. The expression of mRNA for collagen was lower in
the infarct area of FXIII-A⫺/⫺ mice, while the level of
MMP-9 in the myocardial tissue of FXIII-deficient mice was
sevenfold higher than in wild-type mice (268). Substitution
with human pFXIII preparation reversed most, but not all,
changes observed in FXIII-A⫺/⫺ mice. Postmyocardial infarction remodeling was enhanced and left ventricule enddiastolic volume was increased in spite of replenishment
with pFXIII (268). The collagen synthesis remained impaired, and the scar was significantly thinner in these animals. These findings suggest that, although the lack of
pFXIII plays the dominant role in the development of cardiac rupture following ligation-induced myocardial infarction in FXIII-A knockout animals, cFXIII present in the cells
of cardiac tissue also contributes to the pathomechanism.
FXIII-A mRNA and protein have been detected in murine
cardiac tissue, and FXIII-A antigen was localized to spindleand round-shaped cells positive for monocyte/macrophage
marker by immunohistochemical methods (351, 352).
FXIII activity becomes gradually elevated in the myocardial
infarct of wild-type mice with a maximum on day 3 (267,
268). Treatment of these animals by extra pFXIII resulted in
a series of changes. It attenuated left ventricular dilation,
enhanced the recruitment of macrophages in the myocardium, upregulated VEGF, and increased the formation of
new microvessels and the synthesis of collagen (267).
The same survival rate of homozygous and heterozygous
FXIII-A knockout animals after coronary artery ligation is
rather unexpected. However, FXIII has been shown to be
used up during the acute phase of myocardial infarction
(18, 110), and further decrease of the already-low FXIII
level in FXIII-A⫹/⫺ mice could result in a situation similar to
FXIII-A⫺/⫺ animals. Furthermore, diminished FXIII levels
Spontaneous pathological alterations in the cardiac tissue
were also observed in FXIII-A⫺/⫺ mice (352). Interestingly,
these changes seemed to be gender specific. Only ⬃50% of
male FXIII-A⫺/⫺ mice survived 10 mo, while 90% of female
mice of the same genotype remained alive. A quarter of
deceased mice died of severe intra-thoracic hemorrhage,
and large hematomas were found in their hearts. Hemorrhage was also found in the heart of other deceased mice,
and all FXIII-A⫺/⫺ males showed widespread hemosiderin
deposition and fibrosis in the cardiac tissue. Heart structure
and function of FXIII-A knockout mice did not differ significantly from their wild-type counterparts. Fibrosis and
hemosiderin deposition were also present, although to a
much lower extent, in FXIII-B⫺/⫺ males, but not in females.
There is no clear explanation for the gender difference; it
was supposed that the more aggressive fighting behavior of
male animals could be one of the contributing factors.
The above findings suggest that FXIII exerts its cardiac
effect by multiple mechanisms. It upregulates the synthesis
of extracellular matrix components, and in the case of ischemic tissue damage, it enhances microvessel angiogenesis
and modulates the influx of inflammatory cells. Additional
mechanisms might also be revealed in the near future. It also
remains to be seen to what extent the mechanisms operating
in the coronary artery ligation-induced murine infarction
model are valid for human myocardial ischemic damage
caused by atherothombotic events. Acute myocardial infarction developed on the basis of atherosclerosis is a more
complex process in which ischemic cell death is only the
end-stage of a series of events. Clinical findings suggest that
FXIII might contribute to the events leading to coronary
occlusion. Gene expression profiling revealed increased expression of FXIII-A in peripheral blood cells in patients with
coronary artery disease (230), and an elevated FXIII level
was found to be a risk factor of myocardial infarction in
women (40). Detailed discussion of the interrelationship
between FXIII and myocardial infarction is beyond the
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
955
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
FXIII-A⫺/⫺ mice supplemented with human pFXIII and
wild-type mice following myocardial infarction caused by
left coronary artery ligation. FXIII-A⫺/⫺ and FXIII-A⫹/⫺
mice subjected to coronary ligation died within 5 days due
to the rupture of the left ventricle wall. No death occurred
in sham-operated mice, and none of the wild-type coronary
ligated mice died within 5 days (only 40% of them died
within 40 days). Daily intravenous replenishment of FXIIIA⫺/⫺ mice with human pFXIII restored survival to the level
seen in wild-type controls. Immunohistochemistry demonstrated the presence of FXIII-A in the healing infarct of
wild-type and pFXIII supplemented FXIII-A⫺/⫺ mice.
Within the infarct of these animals, considerable FXIIIa
activity was detected by molecular imaging using 111Inlabeled NH2-terminal ␣2-PI dodecapeptide as FXIIIa substrate (156, 268). In FXIII-A⫺/⫺ mice, a reduced inflammatory response was observed at the site of infarct; on day 2
following ligation, the recruitment of neutrophil granulocytes and macrophages to the infarct decreased by 56 and
54%, respectively (268, 269). The diminished inflammatory response is related to the impaired migration of monocytes from their splenic reservoir to the site of the injury
(358). Normally the recruitment of monocytes to the site of
myocardial injury operates through angiotensin II-AT1 receptor signaling (358) that very likely involves the intracellular activation of FXIII (1) (see also sects. VB and VIIB).
The lack of FXIII could abrogate such a mechanism.
MUSZBEK ET AL.
scope of this review; for a most recent overview on this
topic, see Reference 260.
D. Factor XIII in Cartilage and Bone
The function of chondrocyte FXIII seems to change according to the location of the cells and the stage of bone development. In the embryonic avian and mouse growth plate of
long bones, it is not present in cells of proliferative zone and
appears first within the cells of the zone of maturation. Its
expression progressively increases until the prehypertrophic
zone, and in hypertrophic cells, the expression remains at a
high level (280 –282). In articular cartilage, FXIII-A-containing chondrocytes are predominantly localized in the
condylar region of tarsus, suggesting that FXIII is also
needed in regions that will be exposed to mechanical stress
and abrasive forces. On the basis of experiments using cultured hypertrophic chondrocytes transfected with FXIII-Agreen fluorescence protein construct, it was surmised that
cell death and lysis are responsible for the externalization of
FXIII (280). However, this theory could not be verified by
experiments on human embryonic growth plate and articular chondrocytes (165). The incorporation of a lysine-containing rhodamine-labeled tetrapeptide, an amine substrate
of FXIIIa, into intracellular proteins and into extracellular
matrix components indicated that, at least part of chondrocyte FXIII became activated both in the intracellular and
extracellular environment (281, 282). No clear evidence
was provided for its intracellular proteolytic activation in
chondrocytes. However, just like in platelets, nonproteolytic activation of the zymogen, induced by increased Ca2⫹
concentration, might have been sufficient to bring about TG
activity. On the basis of SDS-PAGE analysis, FXIII-A of
avian chondrocyte was mistakenly considered as a monomer (282); it is very likely a homodimer, like cFXIII of other
cell origin.
956
In normal human knee articular cartilage, some expression
of TG-2 and cFXIII was detected in the superficial zone and
in the central (chondrocytic) zone of medial menisci (165).
Markedly upregulated expression of both TGs by enlarged
chondrocytes was observed in the superficial and deep
zones of articular cartilage obtained from patients with severe osteoarthritis. A similar finding was obtained with medial meniscal cartilage. Antibody against FXIII-A had a
greater neutralizing effect on the TG activity of meniscal cell
lysate than anti-TG-2 antibody, which suggests the predominance of cFXIII over TG-2. Because the cell lysates
were not treated with thrombin, the actual cFXIII-to-TG-2
ratio could have been significantly higher than that measured by the authors.
Interleukin-1␤ (IL-1␤) and transforming growth factor-␤
(TGF-␤), whose activities are upregulated in osteoarthritis,
induced increased TG-2 and FXIII-A immunostaining in
knee articular and meniscal cartilage slices carried in organ
culture (165). In cartilage with severe osteoarthritis, IL-1␤
was more effective than TGF-␤. The mechanism of IL-1␤induced TG expression is not clear. IL-1␤ induces nitric
oxide (NO) production, and NO donors also increase TG
expression, although to a lesser extent than IL-1␤. The results suggest that the effect of IL-1␤, at least in part, is
exerted via increased NO generation.
When TC28 cells, an immortalized cell line of human juvenile costal chondrocytes, or cultured meniscal chondrocytes
were transfected with FXIII-A or TG-2, the markedly increased TG activity was accompanied by marked increase
of calcium precipitates in the matrix of cultured cells (165).
This effect of elevated cFXIII and TG-2 did not seem to be
attributable to changes in PPi metabolism or cell apoptosis.
Increased TG activity was not associated with apoptosis of
the TC28 cells or meniscal chondrocytes and failed to in-
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
FXIII-A has been detected by immunohistochemical and
Western blotting methods in hypertrophic chondrocytes of
chicken and mouse embryonic long bone’s growth plate
(281, 282), as well as in avian, porcine, and human articular
chondrocytes (165, 282, 318). Chondrocytes express TG
activity, which is due to TG-2 and cFXIII that coexist in
these cells (165, 280 –282). The highly increased TG activities following treatment with thrombin also prove the presence of functional cFXIII zymogen in these cells (280, 319).
The detection of FXIII-A mRNA indicates that chondrocytes actively synthesize this protein (280, 282, 318).
FXIII-A was detected in the cytosol and membrane fractions of cultured porcine chondrocytes, more in those cells
obtained from old than from young animals (318). FXIII-A
also appeared in the cell culture media and in articular
cartilage vesicles that participate in calcium pyrophosphate
crystal formation. The mechanism of its externalization is
unclear.
More recent experiments reported by Johnson et al. (166)
shed some more light on the mechanism by which FXIII is
involved in the hypertrophic differentiation of chondrocytes. There seems to be an intriguing interplay between
cFXIII and coexisting TG-2. Treatment of human articular
chondrocytes with cFXIII or with TG-2 increased the expression of VEGF and MMP-13 genes, typically expressed
by hypertrophic chondrocytes, and type X collagen synthesis was also enhanced. cFXIII was shown to be a ligand of
␣1-integrin, and extracellularly administered cFXIII induced ␣1␤1-integrin-dependent signaling leading to drastic
translocation of TG-2 from the cytosol to the cell surface
and to consequent p38 MAP kinase and FAK phosphorylation. The latter events have a central role in transducing
maturation to hypertrophy in cultured chondrocytes (241,
373, 387). As it was shown on chondrocytes from FXIII-A
and TG-2 knockout mice, exogenous cFXIII required both
endogenous cFXIII and TG-2 for its effect on chondrocyte
differentiation (166).
MULTIPLE FUNCTIONS OF FACTOR XIII
duce significant changes in the extracellular PPi or nucleoside triphosphate pyrophosphohydrolase or alkaline phosphatase activity. The cross-linking and stabilization of pericellular calcium binding proteins by TGs (12, 13) might
contribute to this process.
The data obtained on MC3T3-E1 cells were confirmed and
extended on mouse bone by the same group (271). The
expression of FXIII-A protein in osteoblasts and osteocytes
of long bones formed by endochondrial ossification and flat
bones formed primarily by intramembraneous ossification
was demonstrated by immunohistochemistry and in situ
hybridization. FXIII-A mRNA was also detected in the osteoblasts of human tibial subchondral bone plates (327).
Interestingly, its expression became significantly upregulated in sclerotic osteoblasts compared with nonsclerotic
cells.
It is rather surprising that FXIII-A in the bone extract and in
MC3T3-E1 cells was found to be a 37 kDa protein (271), as
opposed to the 80 kDa FXIII-A in platelets and monocytes.
The expression of full-length FXIII-A mRNA in MC3T3-E1
cells and in mouse primary osteoblasts, moreover, the presence of 80 kDa protein in forskolin-treated MC3T3-E1
cells, suggest that the 37 kDa band on the Western blot
represents a product of posttranslational proteolytic fragmentation (271). It is a question if the 37 kDa product is the
result of a controlled, limited intracellular proteolysis or it is
produced during the extraction procedure by proteolysis in
the cell lysate. There are several disturbing findings concerning the former possibility. The authors also found, almost exclusively, 37 kDa FXIII-A in mouse hypertrophic
chondrocytes. Similarly, a 37 kDa band together with even
In conclusion, the studies of the last 8 years, described
above, extended FXIII research to a new field, cartilage and
bone physiopathology. Now, it is clear that chondrocytes
and osteoblasts express FXIII-A, and the expression is influenced by their stage of differentiation, location, and agonists of different cellular functions. In these cells, FXIII-A
coexists with another TG, TG-2, and there is an interesting
functional interplay between the two enzymes. cFXIII of
chondrocytes and osteoblasts becomes externalized and activated, although the biochemistry of neither of these events
has been clarified. Experiments with cultured cells suggest
that externalized FXIIIa (and TG-2) is involved in extracellular matrix formation, assembly, and stabilization, which
seem to be important prerequisites for the process of mineralization. Investigations on FXIII-deficient patients and
on FXIII-A knockout mice concerning the structure/development of bone and cartilage has not been carried out in
this respect. The few available clinical data support a pathophysiological role of FXIII related to hypertrophic differentiation of chondrocytes and osteoarthritis. In the coming
years, more extensive research is expected on this area that
might even lay down the scientific foundation of topical
recombinant FXIII therapy in certain pathological conditions or in traumatic bone injuries.
VII. FACTOR XIII AS AN INTRACELLULAR
ENZYME
A. Involvement of Factor XIII in Platelet
Function
Although the presence of FXIII-A in platelets has been
known for half a century, only sporadic pieces of information became available on the intracellular function of cFXIII
in these cells. Platelet FXIII-A used to be considered only as
a source of FXIII-A in pFXIII. However, in the light of
slowly accumulating data on the intracellular activation of
cFXIII and on the cross-linking of proteins within platelets,
this former view could hardly be maintained any more.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
957
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
In addition to chondrocytes, cFXIII and TG-2 were also
detected in cultured MC3T3-E1 (pre)osteoblasts both at
mRNA and protein levels (17). Abundant TG activity was
observed in the lysate of cells undergoing differentiation,
which was highly increased after thrombin treatment, suggesting that cFXIII was the dominant TG in these cells.
Treatment of the cells with ascorbic acid, which stimulated
collagen synthesis and assembly, greatly enhanced the externalization of cFXIII, although the mechanism of externalization has not been explored. The covalent incorporation of monodansyl cadaverine (MDC), a TG amine substrate, into the extracellular matrix proved that at least part
of externalized FXIII existed in active form. Fibronectin
was identified as the major TG substrate matrix component. Abrogation of TG activity by cystamine decreased
matrix accumulation and mineralization induced by the addition of ascorbic acid plus ␤-glycerophosphate (17). Cystamine also blocked cell differentiation program as indicated by the low level of alkaline phosphatase. It is very
likely that the lack of sufficient matrix assembly and alkaline phosphatase level both contributed to the blockade of
mineralization.
lower molecular weight breakdown products are shown on
the Western blot of mouse macrophage lysate. In sharp
contrast to these findings, full-length FXIII-A has been demonstrated in chondrocytes (282) and in mouse macrophages
(320) without significant proteolytic fragmentation. Furthermore, it was also claimed that the 37 kDa FXIII-A fragment is active without proteolysis by thrombin (271), while
in a previous report from the same group thrombin highly
increased the TG activity in MC3T3-E1 cell extracts (17).
Unless a 37 kDa FXIIIa is isolated from these cells and
becomes biochemically well characterized (which sequence
of FXIII-A is represented, what is its activity compared with
the full-length enzyme, etc.), its intracellular existence and
function remain questionable.
MUSZBEK ET AL.
Agonist-induced platelet aggregation in platelet-rich plasma
from patients with FXIII deficiency and from FXIII-A knockout mice was normal (174, 199, 289), and washed FXIIIA-deficient human or mouse platelets showed normal aggregation response to thrombin (199, 262). Although fibrinogen binding to thrombin receptor agonist peptide
(TRAP) activated FXIII-A-deficient platelets was somewhat
decreased (160), this decrease does not seem to be sufficient
to influence the aggregation process. Platelet aggregation is
a relatively fast process; it is over in 5 min, while the crosslinking of platelet proteins is much slower (262). This suggests that cross-linked polymers are required only for the
later phases of platelet activation, e.g., for spreading of
platelets following adhesion or for clot retraction. The
cross-linking of contractile proteins, like actin, myosin, vinculin, and filamin (27, 67, 68, 340), supports such a hypothesis. The role of platelet FXIII in clot retraction is not clear.
Platelet clot retraction of FXIII-deficient patients was found
to be normal (162, 306) or even enhanced (274). In contrast, reduced contractile force was measured in plateletrich plasma clot from a FXIII-deficient patient (57), and the
combination of FXIII-A-deficient plasma with normal
platelets was sufficient to impair isometric contraction of
platelet-rich plasma clot (66). Impaired retraction of clot
formed in platelet-rich plasma of FXIII-A knockout mice
was observed by Kasahara et al. (174). In the latter experiments, the addition of human pFXIII or rFXIII-A2 to the
platelet-rich plasma of FXIII-A knockout mice only partially restored clot retraction, suggesting that cFXIII also
contributes to the process (174). However, in an earlier
report, FXIII-deficient platelets induced normal clot retraction (306). The controversies concerning the involvement of
pFXIII and/or cFXIII in clot retraction might be due to
958
species difference and to the considerable difference between contractile force measurement and clot retraction in
a test tube.
Most recently, an interesting observation was reported concerning the spreading of platelets on fibrinogen-coated surface (160). FXIII-A-deficient platelets demonstrated an altered spreading phenotype compared with normal platelets.
The altered phenotype of FXIII-A-deficient platelets is characterized by higher percentage of platelets emitting filopodial versus lamellipodial extension and by a delay in the
spreading process. The involvement of cFXIII in the spreading process is further supported by experiments in which
MDC induced a dose-dependent increase in the proportion
of control platelets showing filopodial processes (160).
MDC might also influence platelet function by means other
than the inhibition of protein cross-linking by FXIIIa; it can
inhibit calmodulin-dependent mechanisms (76) and modify
proteins at their glutamine residues. However, its ineffectiveness on the spreading of FXIII-A-deficient platelets indicates that the alteration of platelet spreading by this compound was due to the inhibition of FXIIIa-mediated protein
cross-linking.
Kulkarni and Jackson (192) studied the involvement of
FXIII in the late stage of platelet spreading on collagen I
coated surface. Fully spread platelets undergo a specific sequence of morphological changes, which the authors term
sustained calcium-induced platelet (SCIP) morphology. It
started with the marked contraction of lamellipodial membranes leading to microvesiculation and eventual fragmentation of the cells into a larger central membrane structure
surrounded by smaller membrane bodies. SCIPs bind annexin V, indicating the translocation of phosphatidylserine,
the procoagulant phospholipid, to the outer membrane surface. SCIP formation of platelets from a FXIII-A-deficient
patient was decreased, and MDC also attenuated this process. It was claimed that a membrane-impermeable FXIIIa
inhibitor also exerted an inhibitory effect; however, no evidence was given for the extracellular presence of FXIIIa. It
seems that the activation of FXIII and calpain act synergistically in this process.
The above findings indicate that the interaction of FXIII
with cytoskeletal elements plays a role in its intracellular
function. Results published on this area have not been assembled into a comprehensive coherent view; however, important pieces of the puzzle have been revealed. In resting
platelets, FXIII-A is uniformly distributed in the platelet
cytosol (340, 346, 389); however, during activation or
spreading on an adhesive surface, at least part of it becomes
translocated to the cytoskeletal fraction (340, 389). In
platelets activated by thrombin or Ca2⫹ ionophore, cFXIII
translocated to the periphery of the cells within 1 min and
became associated with thin ruffle like extensions and/or
pseudopods, then gradually diffused back into the central
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
Experiments with FXIII-A-deficient platelets provided important data on the role of cFXIII in platelet function. As
discussed in section IIIA, FXIII becomes activated by the
nonproteolytic pathway in human platelets during activation induced by thrombin or Ca2⫹ ionophore (261, 262)
and cross-links platelet proteins into high Mr polymers (67,
68, 128, 262). Although only a smaller portion of platelet
FXIII becomes activated, due to the very high concentration
of cFXIII in these cells, this active fraction is quite sufficient
to bring about the cross-linking reaction (261, 340). It is to
be emphasized that in human platelets FXIII is the only TG.
TG-2 protein and TG-2 mRNA could not be detected in
FXIII-A-deficient or normal human platelets, TG activity is
negligible in FXIII-A-deficient platelets (160, 243, 255),
and no highly cross-linked protein polymers are formed in
these cells (262). There might be a species difference in this
respect; TG activity in the lysate of platelets from FXIIIA⫺/⫺ mice was higher than in platelet lysate from normal
mice, suggesting a compensatory upregulation of another
TG (164). It is clear that no such compensatory mechanism
exists in humans, and protein cross-linking observed in activated platelet is entirely due to the TG activity of FXIIIa.
MULTIPLE FUNCTIONS OF FACTOR XIII
located to the cytoskeleton formed during platelet activation. 4) Intracellular FXIIIa is very likely involved in certain
phases of the platelet spreading process and might be involved in clot retraction, at least in mice.
cFXIII may interact with platelet proteins in two different
ways: 1) its inactive or active form may associate with other
platelet proteins without using them as substrate and move
together during translocation, and 2) FXIIIa can utilize
some proteins as glutamine and/or lysine donor substrates.
Zhu et al. (389) demonstrated the association of FXIII with
the low Mr heat shock protein HSP27 in platelets. In resting
platelets, HSP27 was also found in the cytosol, but upon
activation, it becomes phosphorylated and associated with
the activation-dependent cytoskeleton (240). As opposed to
resting platelets, in adhered platelets HSP27, in colocalization with cFXIII, was preferentially confined to the circumferential region consistent with the outermost extensions of
spread platelets and to the central region corresponding to
condensed granules and contractile elements (389). It is
possible that HSP27 escorts FXIII to specific sites within the
cells, for instance, to the cytoskeleton assembled de novo
during the activation of platelets. Four other proteins, focal
adhesion kinase (FAK), gelsolin, myosin IIA, and thrombospondin I, coimmunoprecipitated with cFXIII from the
platelet lysate and found in colocalization with FXIII-A
both in resting and activated platelets (309). Gelsolin and
myosin IIA are cytoskeletal proteins, while the tyrosine kinase FAK translocates to the cytoskeleton in the second
wave of cytoskeletal rearrangement (285). Direct association between cFXIII and these proteins should be confirmed
in binding experiments with purified proteins.
Most recently, the involvement of cFXIII in certain functions of megakaryocytes has also been reported (231). It
was shown that megakaryocyte adhesion to type I collagen
promoted the spreading of these cells and inhibited proplatelet formation through the release and relocation to the
plasma membrane of cellular fibronectin. cFXIII was also
shown to be translocated to the outer surface of plasma
membrane and cross-linked fibronectin, although the mechanism of its translocation was not revealed. The inhibition
of FXIIIa significantly decreased megakaryocyte spreading
on type I collagen and reduced actin stress fiber formation
and the assembly of fibronectin. These findings suggest that
FXIII participates in the mechanism by which type I collagen, in the osteoblastic niche, suppresses megakaryocyte
maturation and prevents premature platelet release.
The cross-linking of two cytoskeletal elements, filamin and
vinculin, during platelet activation has been analyzed in
detail (340). A high portion of filamin was cross-linked to at
least four different nonidentified platelet proteins in thrombin-activated platelets within 10 min. Blocking TG activity
by the -SH reagent iodoacetamide completely abolished filamin cross-linking. Inhibition of fibrinogen binding to
platelets and, consequently, platelet aggregation by the peptide Gly-Pro-Arg-Pro decreased the cross-linking of filamin,
suggesting that it is an aggregation-induced process. The
cross-linking of vinculin was less effective, and it was a
platelet aggregation-dependent slower process.
In summary, although the exact biochemical mechanism by
which FXIII participates in platelet function still needs clarification, the available data allow the following conclusions: 1) a part of platelet FXIII becomes transformed into
an active TG during platelet activation by the nonproteolytic activation mechanism. 2) Activated FXIII, the only TG
in human platelets, cross-links platelet proteins in the later
phases of platelet activation. 3) A few cytoskeletal proteins
have been identified as substrates for FXIIIa in the crosslinking reaction, and at least part of FXIII becomes trans-
B. Involvement of Factor XIII in Monocyte/
Macrophage Function
Since the discovery of FXIII-A in monocytes and macrophages in 1985 (8, 130, 257), its expression in various
monocyte-derived mobile and fixed macrophages has been
intensively studied and turned out to be a marker reaction
for various subsets of these cells in physiological and pathological settings (6, 304). The possible coexistence of cFXIII
with TG-2 in monocytes and macrophages is a somewhat
debated question. We could not measure any detectable TG
activity in monocytes from four patients with severe
FXIII-A deficiency (255), and only negligible TG activity
was measured in such monocytes by the highly sensitive
[3H]putrescine incorporation assay (72). No TG-2 was
found in normal peripheral blood monocytes by Seiving et
al. (338), and only minute TG activity was measured in
nonactivated human monocyte cell lysate without thrombin activation (239, 242, 254). In the latter studies, Western
blotting also detected a small amount of TG-2. The amount
of TG-2 was much less than that of FXIII-A in the antigenspecific immunoprecipitates from nonactivated monocytes
(16). Most recently, similar FXIII-A and TG-2 mRNA levels
were measured in human monocytes; however, in this
study, no attempt was made to measure the translated protein levels (161). In conclusion, overwhelming pieces of evidence suggest that in resting nonactivated monocytes,
cFXIII is the major TG, and the amount of TG-2, if any, is
not significant. However, during differentiation into macrophages and following activation by various agents, major
changes occur in the intracellular level of both TGs.
In two early studies, opposing results were obtained on the
changes of cFXIII in monocyte culture. A three- to fourfold
increased FXIII-A activity and antigen levels were measured
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
959
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
area (340). Cytochalasin D that inhibits the polymerization
of actin in intact cells abolished the translocation of cFXIII
to the platelet periphery. Part of cytoskeleton-associated
cFXIII existed in active form.
MUSZBEK ET AL.
on day 8 (72), while decreasing cFXIII with increasing TG-2
was reported in long-term culture of monocytes (338).
More recently, a significant elevation of both FXIII-A
mRNA and protein level was observed during the course of
monocyte/macrophage differentiation in cell culture that
peaked on day 3 (329). It is to be noted that the expression
of TG-2 also becomes upregulated during culturing of
monocytes (239, 242, 254). It is interesting that in monocytic leukemia cells, cFXIII level was also increased compared with normal peripheral blood monocytes (172).
IL-4 is also involved in the in vitro differentiation of monocytes into antigen-presenting dendritic cells. Induction of
such differentiation by granulocyte/macrophage-colony
stimulating factor (GM-CSF) plus IL-4 resulted in very high
intracellular levels of FXIII-A mRNA and protein (15, 161,
362). If dendritic cells were treated by BCG, stimulating the
classical pathway, FXIII-A expression returned to the basal
level within 2 days (362). Subsets of dendritic cells present
in tissues, like subsets of dermal dendrocytes, also express
FXIII-A. The characterization of these subsets is reviewed in
Reference 304.
cFXIII of perivascular macrophages has been implicated in
the flow-dependent remodeling of small arteries in mice (30,
388). In this process, the perivascular accumulation of
FXIII-A-positive macrophages was dependent on CXCchemokine receptor 3 (CXCR3) signaling by the CXCR3
960
Considering the high number of reports on the expression
of FXIII-A in monocytes, macrophages, histiocytes, and
dendritic cells, it is surprising that only a few studies have
concerned with the intracellular function of cFXIII in these
cells. Monocytes from FXIII-deficient patients, which are
exempt of cFXIII, showed an impaired capacity of Fc␥,
complement, and lectin-like receptor-mediated phagocytosis (329). In cell culture, the expression of FXIII-A and the
phagocytosing capacity of monocytes/macrophages increased in parallel, and the competitive FXIIIa substrate,
MDC, significantly inhibited receptor-mediated phagocytosis by these cells. DD cells (a human myelomonocytic cell
line) were incapable of phagocytosis and did not express
FXIII-A antigen. Partial differentiation of the cells, achieved
by phorbol ester treatment, resulted in increased phagocytosis in parallel with the expression of FXIII-A (177). Jayo
et al. (161) investigated the role of cFXIII in the locomotion
of monocyte-derived dendritic cells. MDC inhibited both
basal and chemokine (C-C motif) ligand 19 (CCL19)-induced migration of mature dendritic cells, and FXIII-Adeficient dendritic cells showed a reduced chemotactic response to CCL19. CHO cells overexpressing FXIII-A
showed enhanced mobility both in trans-well and scratch
wound migration assays and displayed membrane blebs
and dynamic cell protrusions involved in cell movement. As
both phagocytosis and cell locomotion heavily involve intracellular contractile elements, the interaction of cFXIII or
its activated form with cytoskeletal components in monocytes/macrophages seems rather feasible. However, this theory still lacks experimental support.
An interesting aspect of the intracellular function of cFXIII
in monocytes was reported by AbdAlla et al. (1). In monocytes from hypertensive patients, covalently cross-linked
angiotensin receptor 1 (AT1) dimers were detected; this
dimerization was carried out by nonproteolytically activated cFXIII. The covalent cross-linking of AT1 could be
reproduced by the combined addition of angiotensin II and
ionomycin, which activated AT1 and, by increasing the intracellular Ca2⫹ concentration, cFXIII. No cross-linking of
AT1 was observed in monocytes from FXIII-A-deficient patients, and it was also abolished by angiotensin convertase
inhibitor. Cross-linked AT1 dimers displayed enhanced
G␣q/11-stimulated signaling, increased internalization, and
desensitization. Gln315 in the cytoplasmic side of AT1 provided the Gln substrate for the FXIIIa-catalyzed cross-linking. The appearance of cross-linked AT1 dimers correlated
with the enhanced angiotensin II-dependent adhesion of
monocytes to endothelial cells and endothelial dysfunction.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
Torocsik et al. (362) investigated the changes of FXIII-A
expression after activation of monocytes through the classical or alternative pathway. Induction of the classical activation pathway by interferon (IFN)-␥ or Mycobacterium
bovis BCG downregulated FXIII-A mRNA and protein expression in cultured macrophages, while activation by IL-4
through the alternative pathway exerted the opposite effect.
In the latter case, on days 4 and 5, the increase of FXIII-A
mRNA level was more than 40-fold, and FXIII-A protein
level became elevated 11-fold on day 5. The expression of
TG-2 did not change in these conditions. Similar upregulation of FXIII-A mRNA by IL-4, and to a lesser extent by
IL-13, was observed by another group, as well (59), while
priming monocytes by lipopolysaccharide decreased
FXIII-A expression (290). In another study, the effect of
IL-4 on the expression of FXIII-A at the mRNA and protein
level was highly increased by the addition of dexamethasone (120). Results of in vitro investigations were also supported by the detection of FXIII-A in tumor-associated
macrophages that are considered as alternatively activated,
and its absence in macrophages of tuberculotic granuloma
that go through the classical activation pathway (362). The
expression of FXIIIa substrate fibronectin is also augmented by IL-4/IL-13. The parallel upregulation of cFXIII
expression and tissue inhibitor of metalloproteinase 3
(TIMP3) could protect the deposited extracellular matrix
protein from degradation (59).
ligand, interferon inducible protein-10 (IP-10). This mechanism also contributed to the highly increased production
of FXIII-A in CRCR3-positive macrophages (388). However, it is not clear how elevated cFXIII in perivascular
macrophages is involved in vascular remodeling and if such
a mechanism also operates in humans.
MULTIPLE FUNCTIONS OF FACTOR XIII
The inhibition of angiotensin II generation or intracellular
FXIII activity suppressed the appearance of cross-linked
AT1 receptors and symptoms of atherosclerosis in apoEdeficient mice. Follow-up studies confirming these exciting
findings are awaited.
There is not much information on the release or surface
expression of monocyte/macrophage cFXIII. Although
there is a general agreement on the cytoplasmic localization
of FXIII-A in these cells, a few studies indicated that macrophages in cell culture translocate it to the surface (16, 72,
190). Increased surface staining of U937 cells (a human
promonocytic tumor line) for FXIII-A was observed after
incubation with phorbol myristate acetate, lipopolysaccharide, and IFN-␥ (190). In a single study, FXIII-A was detected in the culture medium of dendritic cells differentiated
from peripheral blood monocytes (15). However, in these
studies, dying or dead cells were not excluded as the source
of cFXIII appearing on the cell surface or in the medium.
Although the surface expression or release of cFXIII from
macrophages, especially from resident macrophages and its
involvement in organization and remodeling of extracellular matrix by the cross-linking matrix proteins is quite plausible, this hypothesis has not been supported by experimental proof. As mentioned in section IIC, there is evidence for
cFXIII entering the alternative secretory pathway in human
macrophages (75), but its actual secretion has not been
proven.
In addition to the possible translocation of cFXIII from the
cytoplasm to the cell surface, another type of cFXIII translocation has also been revealed in macrophages. The transient appearance of cFXIII in the nuclei was demonstrated
by confocal laser scanning microscopy, Western blotting,
and immunoelectron microscopy in the early phase of
monocyte/macrophage differentiation (7). On day 2 of culturing, 90% of the cells expressed FXIII-A in the nucleus in
association with electrodense areas, but this phenomenon
was not observed on any other day. cFXIII in the nucleus
somehow became activated, and its TG activity was demonstrated by the incorporation of MDC into nuclear proteins. Nuclear localization of TG-2 has also been observed
(208, 291, 295), but the function of neither TG has been
revealed in the nuclear environment. Follow-up studies on
VIII. SUMMARY, CONCLUSIONS, AND
PERSPECTIVES
pFXIII is a tetrameric zymogen protransglutaminase, which
consists of two potentially active A and two inhibitory/
carrier B subunits (FXIII-A2B2). In the plasma it is bound to
the ␥=-chain of fibrinogen through FXIII-B subunits. The
synthesis of the two subunits occurs in different cells, and
they form complex in the plasma. There is little doubt that
the major part of FXIII-A is synthesized by cells of bone
marrow origin, while FXIII-B is produced and secreted by
hepatocytes. However, it is still unclear what is the contribution of megakaryocytes and monocytes/macrophages to
FXIII-A present in pFXIII. The secretory pathway by which
FXIII-A is released from the cells also remains to be elucidated, although most recently evidence has been provided
on its entering into the nonclassical secretory pathway in
macrophages. FXIII-B is in excess in the plasma; ⬃50% of
it circulates in free, uncomplexed form. The dimer of
FXIII-A, but not FXIII-B, is also present in cells (platelets,
monocytes, macrophages, chondrocytes, osteoblasts, osteocytes). The three-dimensional structure of FXIII-A2 has
been revealed by X-ray crystallography, and four major
structural domains (␤-sandwich domain, central core domain, and two ␤-barrel domains) and AP-FXIII were identified. The three-dimensional structure of FXIII-B is not
known; it contains 10 sushi domains each held together by
two pairs of disulfide bonds.
pFXIII becomes activated in the final phase of coagulation
cascade by thrombin and Ca2⫹. Thrombin cleaves off APFXIII from FXIII-A, which, as demonstrated most recently,
dissociates from the truncated molecule (FXIII-A=). The
newly formed fibrin ⬃100-fold accelerates this process. In
the presence of Ca2⫹, FXIII-B leaves FXIII-A=2, and the
latter assumes an enzymatically active configuration (FXIIIA*2). In plasma, the activation of FXIII occurs on the surface of fibrin, and FXIIIa remains associated with it, while
FXIII-B is released into the serum. The release of AP-FXIII
is not a prerequisite of FXIII activation. At very high unphysiological Ca2⫹ concentration (ⱖ100 mM), FXIII-B dissociates from the nontruncated FXIII-A2, which then becomes enzymatically active (FXIII-A°2). The main role of
the truncation of FXIII-A is to bring down the Ca2⫹ requirement for the dissociation step into physiological range. In
the absence of FXIII-B, cFXIII can be transformed into
FXIII-A°2 at low Ca2⫹ concentration. The main role of
FXIII-B seems to prevent FXIII-A2 from going through a
slow progressive activation in plasmatic environment. Several pieces of evidence suggest that the nonproteolytic pathway is the main form of cFXIII activation in platelets and in
monocytes/macrophages. No major conformational change
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
961
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
In a most recent paper it was demonstrated that cFXIII is
involved in the regulation of gene expression in alternatively activated human macrophages (363). In microarray
experiments, a significant difference was demonstrated in
the IL-4-sensitive gene expression profile of cultured macrophages derived from FXIII-deficient and nondeficient individuals. The most prominent differences were related to
immune functions and wound response, suggesting that the
functional impairment of macrophages at the level of gene
expression regulation plays a role in the wound healing
defect of FXIII-deficient patients.
the involvement of cFXIII in nuclear processes are required
to extend and clarify the significance of this intriguing finding.
MUSZBEK ET AL.
was detected in the structure of FXIII-A2 during transformation into a TG by X-ray crystallography; however, biochemical evidence and the analogy with the activation of
TG-2 indicate that in solution the activation process induces considerable structural changes.
The downregulation of FXIIIa has been explored only most
recently. Polymorphonuclear granulocytes incorporated in
the clot become activated and release several proteolytic
enzymes into their surroundings. These proteases, primarily
elastase, break down FXIIIa within the plasma clot, even in
the presence of serine protease plasma inhibitors. The kinetics of FXIIIa inactivation and fibrinolytic effect of granulocyte proteases ensure the formation of stable fibrin clot,
but prevent the production of over-cross-linked plasma
clot, which would be difficult to eliminate, when it is not
needed any more. Further research on this important field
would broaden our knowledge on what is happening with
FXIIIa in the plasma clot and in the thrombus.
The main function of FXIIIa in the plasma is to cross-link
fibrin chains and ␣2-PI to fibrin. Fibrin ␥-chains are crosslinked into dimers at a very early stage of clotting, while
fibrin ␣-chains form high-molecular-weight polymers at a
much slower rate. The cross-linking of ␣2-PI to fibrin
␣-chains is also rather fast, and the heterodimer becomes
progressively incorporated into the highly cross-linked
␣-chain polymers. Fibrin cross-linking stabilizes fibrin and
makes it more resistant against shear stress. Both fibrin
cross-linking and the covalent incorporation of ␣2-PI into
962
The involvement of FXIII in wound healing has been supported by sporadic clinical data from studies with FXIII-Adeficient patients. Its role in wound healing is now well
proven by the impaired wound healing of FXIII-A knockout
mice. The promotion of wound healing by FXIII is a complex process; it involves its effects on fibroblasts, macrophages, extracellular matrix, and angiogenesis. FXIIIa enhanced the migration of fibroblasts and macrophages, increased the rate of fibroblasts proliferation, and decreased
their apoptosis. Cross-linking of extracellular matrix proteins might modulate these direct effects. The FXIIIa-induced biochemical mechanisms in these events have not
been enlightened, yet. The downregulation of TSP-1 has
been shown to contribute to the effect of FXIIIa on monocytes.
In the last couple of years, the proangiogenic effect of FXIII
has been explored to a significant extent. Although nonactivated FXIII can bind to endothelial cells, the proangiogenic
effect is exerted only by the active form. FXIIIa enhances endothelial cell migration and proliferation and suppresses their
apoptosis. The role of FXIII in angiogenesis has also been
verified in several animal models. The effect of FXIIIa on endothelial cells involves binding to ␣v␤3 receptor, the interaction of this receptor with VEGFR-2, VEGFR-2 phosphorylation, phosphorylation and activation of Akt and ERKs, the
upregulation of the transcription factors c-Jun and Egr-1, and
the downregulation of TSP-1. Further exploration of this
mechanism is expected, and the proangiogenic effect of FXIIIa
or its inhibition could have a prospect of clinical utilization.
Clinical findings and experiments with FXIII-A knockout
mice prove, without any doubt, that pFXIII is essential for
maintaining pregnancy. It is surprising that, in spite of the
importance of these observations, hardly any studies explored the mechanism by which pFXIII prevents spontaneous abortions. As fibrinogen deficiency also leads to early
pregnancy loss, the cooperation of the two clotting factors
seems essential. The cross-linking of fibrinoid components
in the placenta (fibrin, fibronectin) is very likely important
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
FXIIIa, like other TGs, catalyzes a double displacement acyl
transfer reaction, in which the acyl donor peptide-bound
glutamine residue forms thioester with the active site
Cys314 of the core domain, and ammonia is released. In the
second step, an acyl acceptor primary amine is cross-linked
to the glutamine residue through isopeptide bond, and the
active site cysteine becomes deacylated. If the acyl acceptor
amine is provided by a peptide-bound lysine residue, two
peptide chains become covalently cross-linked. The glutamine substrate specificity of FXIIIa is rather restrictive,
only less than two dozen proteins have been identified as
FXIIIa substrates. However, the amino acid sequences
around the substrate glutamine site show considerable variation, and no consensus sequence could be derived from the
known primary structures. The structural requirement of
the interaction with FXIIIa was studied only in the case of
two peptide substrates. One can conclude from the NMR
studies and kinetic experiments with chemical mutant peptides that, in addition to the active site, binding of the substrate peptide to a second interacting site, involving hydrophobic interaction, on the surface of FXIII-A*2 is also a
prerequisite of the enzymatic action. Experiments with a
series of individual glutamine donor peptide substrate are to
be carried out to clarify the structural elements required for
effective enzyme-substrate interaction.
fibrin protect fibrin against fibrinolysis. The cross-linking of
␣2-PI to the fibrin network seems to play the dominant role
in preventing newly formed fibrin from the prompt elimination by plasmin, while fibrin ␣-chain cross-linking could
contribute to the increased resistance of mature thrombi to
fibrinolysis. pFXIII is essential for maintaining hemostasis;
its deficiency results in very severe bleeding diathesis.
pFXIII is now considered as a key player in the process of
fibrinolysis, which exerts its effect through the two mechanisms discussed above. In addition to ␣2-PI, FXIIIa also
cross-links other components of fibrinolysis, like PAI-2,
TAFI, and plasminogen, to fibrin or fibronectin; however,
the significance of these cross-linking reactions remains to
be explored. Further intensive research is expected on this
field in the near future.
MULTIPLE FUNCTIONS OF FACTOR XIII
for the sealing/anchoring effect and barrier function of the
Nitabuch’s layer. The proangiogenic effect of FXIIIa might
contribute to normal placenta development. Elaboration of
these issues could lead to highly important pieces of information and to better understanding of the mechanisms required for maintaining pregnancy.
The active form of FXIII-A2 is also an intracellular enzyme.
Although its presence in platelets is known for more than a
half century and their existence in monocytes/macrophages
for a quarter of century, its intracellular function started to
be addressed only most recently. During platelet activation
part of its cFXIII becomes activated by the nonproteolytic
mechanism in the later phase of the activation process. During this process, part of FXIII becomes translocated to the
cytoskeleton and cross-links platelet proteins, particularly
cytoskeletal proteins. Activated FXIII is involved in certain
phases of platelet spreading. Information on the involvement of cFXIII in clot retraction is controversial. It is not
clear if partial contribution of cFXIII to the clot retraction
in mice could be extrapolated to humans.
Intriguing novel findings were reported on monocyte/macrophage cFXIII in the last couple of years. The expression of
cFXIII in monocytes activated by the alternative pathway
was highly upregulated, while activation through the classical pathway resulted in the opposite effect. The results
obtained with cultured monocytes were also confirmed
with in situ activated tissue macrophages. The transient
appearance of cFXIII in the nucleus on day 2 of monocyte
culturing is an intriguing finding that warrants further investigation. Differentiation of monocytes into dendritic
cells resulted in very high intracellular level of cFXIII. Impaired receptor-mediated phagocytosis of monocytes from
FXIII-deficient patients suggests the involvement of cFXIII
in the biochemical mechanism of this process. cFXIII has
ACKNOWLEDGMENTS
Address for reprint requests and other correspondence: L.
Muszbek, Clinical Research Center, Univ. of Debrecen,
Medical and Health Science Center, 98 Nagyerdei krt., Debrecen 4032, Hungary (e-mail: [email protected]).
GRANTS
This review was made possible through research funding
from the Hungarian National Research Fund (OTKANKTH CNK 80776), the Hungarian Academy of Sciences
(TKI 227), and the Hungarian Ministry of Health (ETT).
The work was also supported by the TÁMOP 4.2.1./B-09/
1/KONV-2010-0007 project. The project is implemented
through the New Hungary Development Plan, cofinanced
by the European Social Fund.
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared
by the authors.
REFERENCES
1. AbdAlla S, Lother H, Langer A, el Faramawy Y, Quitterer U. Factor XIIIA transglutaminase crosslinks AT1 receptor dimers of monocytes at the onset of atherosclerosis. Cell 119:343–354, 2004.
2. Achyuthan KE, Mary A, Greenberg CS. The binding sites on fibrin(ogen) for guinea pig
liver transglutaminase are similar to those of blood coagulation factor XIII. Characterization of the binding of liver transglutaminase to fibrin. J Biol Chem 263:14296 –
14301, 1988.
3. Adamson R. Role of macrophages in normal wound healing: an overview. J Wound
Care 18:349 –351, 2009.
4. Adany R. Intracellular factor XIII: cellular distribution of factor XIII subunit a in humans. Semin Thromb Hemost 22:399 – 408, 1996.
5. Adany R, Antal M. Three different cell types can synthesize factor XIII subunit A in the
human liver. Thromb Haemost 76:74 –79, 1996.
6. Adany R, Bardos H. Factor XIII subunit A as an intracellular transglutaminase. Cell Mol
Life Sci 60:1049 –1060, 2003.
7. Adany R, Bardos H, Antal M, Modis L, Sarvary A, Szucs S, Balogh I. Factor XIII of blood
coagulation as a nuclear crosslinking enzyme. Thromb Haemost 85:845– 851, 2001.
8. Adany R, Belkin A, Vasilevskaya T, Muszbek L. Identification of blood coagulation
factor XIII in human peritoneal macrophages. Eur J Cell Biol 38:171–173, 1985.
9. Adany R, Glukhova MA, Kabakov AY, Muszbek L. Characterisation of connective
tissue cells containing factor XIII subunit a. J Clin Pathol 41:49 –56, 1988.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
963
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
In addition to its role in hemostasis, wound healing/angiogenesis, and maintaining pregnancy, studies in the last decade drew attention to other less evident, less expected, and
less explored effects of FXIII. A few reports indicated that
FXIIIa reduces vascular permeability, but the mechanism is
unknown. Experiments with FXIII-A-deficient mice suggested a protective effect of FXIII against spontaneous cardiac rupture and against cardiac rupture following ligationinduced myocardial infarction. Studies on FXIII in cartilage
and bone extended FXIII research to a new field. It is clear
that both chondrocytes and osteoblasts express FXIII-A,
and its expression is related to the stage of differentiation
and the location of these cells. By an unknown mechanism,
FXIII becomes externalized and activated. Experiments
with cultured cells suggest that it is involved in extracellular
matrix assembly, stabilization, and mineralization. Although some of the reported findings need confirmation,
research on these novel functions of FXIII is highly promising.
also been implicated in the locomotion of dendritic cells.
The nonproteolytic activation of cFXIII in monocytes of
hypertensive patients and the cross-linking of AT1 resulted
in enhanced signaling and adhesion to endothelial cells. The
role of this mechanism in atherosclerosis was suggested by
experiments with apoE-deficient mice. There is no doubt
that further studies on the intracellular function of cFXIII
would lead to exciting new findings and provide new data
on the role of this intracellular TG.
MUSZBEK ET AL.
10. Adany R, Kiss A, Muszbek L. Factor XIII: a marker of mono- and megakaryocytopoiesis. Br J Haematol 67:167–172, 1987.
11. Adany R, Muszbek L. Immunohistochemical detection of factor XIII subunit a in
histiocytes of human uterus. Histochemistry 91:169 –174, 1989.
31. Balogh I, Szoke G, Karpati L, Wartiovaara U, Katona E, Komaromi I, Haramura G,
Pfliegler G, Mikkola H, Muszbek L. Val34Leu polymorphism of plasma factor XIII:
biochemistry and epidemiology in familial thrombophilia. Blood 96:2479 –2486, 2000.
32. Bangert K, Johnsen AH, Christensen U, Thorsen S. Different N-terminal forms of
alpha 2-plasmin inhibitor in human plasma. Biochem J 291:623– 625, 1993.
12. Aeschlimann D, Kaupp O, Paulsson M. Transglutaminase-catalyzed matrix crosslinking in differentiating cartilage: identification of osteonectin as a major glutaminyl
substrate. J Cell Biol 129:881– 892, 1995.
33. Barkan G. Zur Frage der Reversibilitat der Fibringerinnung II. Biochem Ztschr 139:
291–301, 1923.
13. Aeschlimann D, Mosher D, Paulsson M. Tissue transglutaminase and factor XIII in
cartilage and bone remodeling. Semin Thromb Hemost 22:437– 443, 1996.
34. Barry EL, Mosher DF. Binding and degradation of blood coagulation factor XIII by
cultured fibroblasts. J Biol Chem 265:9302–9307, 1990.
14. Ahvazi B, Kim HC, Kee SH, Nemes Z, Steinert PM. Three-dimensional structure of
the human transglutaminase 3 enzyme: binding of calcium ions changes structure for
activation. EMBO J 21:2055–2067, 2002.
35. Barry EL, Mosher DF. Factor XIII cross-linking of fibronectin at cellular matrix assembly sites. J Biol Chem 263:10464 –10469, 1988.
16. Akimov SS, Belkin AM. Cell surface tissue transglutaminase is involved in adhesion and
migration of monocytic cells on fibronectin. Blood 98:1567–1576, 2001.
17. Al-Jallad HF, Nakano Y, Chen JL, McMillan E, Lefebvre C, Kaartinen MT. Transglutaminase activity regulates osteoblast differentiation and matrix mineralization in
MC3T3–E1 osteoblast cultures. Matrix Biol 25:135–148, 2006.
18. Alkjaersig N, Fletcher AP, Lewis M, Ittyerah R. Reduction of coagulation factor XIII
concentration in patients with myocardial infarction, cerebral infarction, and other
thromboembolic disorders. Thromb Haemost 38:863– 873, 1977.
19. Ambrus A, Banyai I, Weiss MS, Hilgenfeld R, Keresztessy Z, Muszbek L, Fesus L.
Calcium binding of transglutaminases: a 43Ca NMR study combined with surface
polarity analysis. J Biomol Struct Dyn 19:59 –74, 2001.
20. Andersen MD, Kjalke M, Bang S, Lautrup-Larsen I, Becker P, Andersen AS, Olsen OH,
Stennicke HR. Coagulation factor XIII variants with altered thrombin activation rates.
Biol Chem 390:1279 –1283, 2009.
21. Ando Y, Imamura S, Yamagata Y, Kitahara A, Saji H, Murachi T, Kannagi R. Platelet
factor XIII is activated by calpain. Biochem Biophys Res Commun 144:484 – 490, 1987.
22. Anwar R, Gallivan L, Edmonds SD, Markham AF. Genotype/phenotype correlations
for coagulation factor XIII: specific normal polymorphisms are associated with high or
low factor XIII specific activity. Blood 93:897–905, 1999.
23. Ariens RA, Philippou H, Nagaswami C, Weisel JW, Lane DA, Grant PJ. The factor XIII
V34L polymorphism accelerates thrombin activation of factor XIII and affects crosslinked fibrin structure. Blood 96:988 –995, 2000.
24. Asahina T, Kobayashi T, Okada Y, Goto J, Terao T. Maternal blood coagulation factor
XIII is associated with the development of cytotrophoblastic shell. Placenta 21:388 –
393, 2000.
25. Asahina T, Kobayashi T, Okada Y, Itoh M, Yamashita M, Inamato Y, Terao T. Studies
on the role of adhesive proteins in maintaining pregnancy. Horm Res 50 Suppl 2:37– 45,
1998.
26. Asahina T, Kobayashi T, Takeuchi K, Kanayama N. Congenital blood coagulation
factor XIII deficiency and successful deliveries: a review of the literature. Obstet
Gynecol Surv 62:255–260, 2007.
36. Beck E, Duckert F, Ernst M. The influence of fibrin stabilizing factor on the growth of
fibroblasts in vitro and wound healing. Thromb Diath Haemorrh 6:485– 491, 1961.
37. Bendixen E, Borth W, Harpel PC. Transglutaminases catalyze cross-linking of plasminogen to fibronectin and human endothelial cells. J Biol Chem 268:21962–21967,
1993.
38. Bendixen E, Harpel PC, Sottrup-Jensen L. Location of the major epsilon-(gammaglutamyl)lysyl cross-linking site in transglutaminase-modified human plasminogen. J
Biol Chem 270:17929 –17933, 1995.
39. Beninati S, Nicolini L, Jakus J, Passeggio A, Abbruzzese A. Identification of a substrate
site for transglutaminases on the human protein synthesis initiation factor 5A. Biochem
J 305:725–728, 1995.
40. Bereczky Z, Balogh E, Katona E, Czuriga I, Edes I, Muszbek L. Elevated factor XIII
level and the risk of myocardial infarction in women. Haematologica 92:287–288,
2007.
41. Biland L, Duckert F. Coagulation factors of the newborn and his mother. Thromb Diath
Haemorrh 29:644 – 651, 1973.
42. Bishop PD, Teller DC, Smith RA, Lasser GW, Gilbert T, Seale RL. Expression, purification, and characterization of human factor XIII in Saccharomyces cerevisiae. Biochemistry 29:1861–1869, 1990.
43. Board PG. Genetic polymorphism of the B subunit of human coagulation factor XIII.
Am J Hum Genet 32:348 –353, 1980.
44. Board PG, Losowsky MS, Miloszewski KJ. Factor XIII: inherited and acquired deficiency. Blood Rev 7:229 –242, 1993.
45. Board PG, Pierce K, Coggan M. Expression of functional coagulation factor XIII in
Escherichia coli. Thromb Haemost 63:235–240, 1990.
46. Board PG, Webb GC, McKee J, Ichinose A. Localization of the coagulation factor XIII
A subunit gene (F13A) to chromosome bands 6p24 —-p25. Cytogenet Cell Genet
48:25–27, 1988.
47. Boda Z, Pfliegler G, Muszbek L, Toth A, Adany R, Harsfalvi J, Papp Z, Tornai I, Rak K.
Congenital factor XIII deficiency with multiple benign breast tumours and successful
pregnancy with substitutive therapy. A case report. Haemostasis 19:348 –352, 1989.
48. Bohn H, Schwick HG. Isolation and characterization of a fibrin-stabilizing factor from
human placenta. Arzneimittelforschung 21:1432–1439, 1971.
49. Borth W, Chang V, Bishop P, Harpel PC. Lipoprotein (a) is a substrate for factor XIIIa
and tissue transglutaminase. J Biol Chem 266:18149 –18153, 1991.
27. Asijee GM, Muszbek L, Kappelmayer J, Polgar J, Horvath A, Sturk A. Platelet vinculin:
a substrate of activated factor XIII. Biochim Biophys Acta 954:303–308, 1988.
50. Bottenus RE, Ichinose A, Davie EW. Nucleotide sequence of the gene for the b
subunit of human factor XIII. Biochemistry 29:11195–11209, 1990.
28. Bagoly Z, Fazakas F, Komaromi I, Haramura G, Toth E, Muszbek L. Cleavage of factor
XIII by human neutrophil elastase results in a novel active truncated form of factor XIII
A subunit. Thromb Haemost 99:668 – 674, 2008.
51. Broekman MJ, Handin RI, Cohen P. Distribution of fibrinogen, platelet factors 4 and
XIII in subcellular fractions of human platelets. Br J Haematol 31:51–55, 1975.
29. Bagoly Z, Haramura G, Muszbek L. Down-regulation of activated factor XIII by polymorphonuclear granulocyte proteases within fibrin clot. Thromb Haemost 98:359 –
367, 2007.
30. Bakker EN, Pistea A, Spaan JA, Rolf T, de Vries CJ, van Rooijen N, Candi E, VanBavel
E. Flow-dependent remodeling of small arteries in mice deficient for tissue-type
transglutaminase: possible compensation by macrophage-derived factor XIII. Circ Res
99:86 –92, 2006.
964
52. Broker M, Bauml O. New expression vectors for the fission yeast Schizosaccharomyces pombe. FEBS Lett 248:105–110, 1989.
53. Brown LF, Lanir N, McDonagh J, Tognazzi K, Dvorak AM, Dvorak HF. Fibroblast
migration in fibrin gel matrices. Am J Pathol 142:273–283, 1993.
54. Bruhn HD, Zurborn KH. Influences of clotting factors (thrombin, factor XIII) and of
fibronectin on the growth of tumor cells and leukemic cells in vitro. Blut 46:85– 88,
1983.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
15. Akagi A, Tajima S, Ishibashi A, Matsubara Y, Takehana M, Kobayashi S, Yamaguchi N.
Type XVI collagen is expressed in factor XIIIa⫹ monocyte-derived dermal dendrocytes and constitutes a potential substrate for factor XIIIa. J Invest Dermatol 118:267–
274, 2002.
MULTIPLE FUNCTIONS OF FACTOR XIII
55. Buluk K. An unknown action of blood platelets; preliminary communication. Pol Tyg
Lek 10:191, 1955.
56. Carpenter SL, Mathew P. Alpha2-antiplasmin and its deficiency: fibrinolysis out of
balance. Haemophilia 14:1250 –1254, 2008.
57. Carr ME ,Jr, Carr SL, Tildon T, Fisher LM, Martin EJ. Batroxobin-induced clots exhibit
delayed and reduced platelet contractile force in some patients with clotting factor
deficiencies. J Thromb Haemost 1:243–249, 2003.
58. Carrell NA, Erickson HP, McDonagh J. Electron microscopy and hydrodynamic properties of factor XIII subunits. J Biol Chem 264:551–556, 1989.
59. Chaitidis P, O’Donnell V, Kuban RJ, Bermudez-Fajardo A, Ungethuem U, Kuhn H.
Gene expression alterations of human peripheral blood monocytes induced by medium-term treatment with the TH2-cytokines interleukin-4 and -13. Cytokine 30:
366 –377, 2005.
60. Chen R, Doolittle RF. Cross-linking sites in human and bovine fibrin. Biochemistry
10:4487– 4491, 1971.
62. Claes L, Burri C, Gerngross H, Mutschler W. Bone healing stimulated by plasma factor
XIII. Osteotomy experiments in sheep. Acta Orthop Scand 56:57– 62, 1985.
63. Cleary DB, Doiphode PG, Sabo TM, Maurer MC. A non-reactive glutamine residue of
alpha2-antiplasmin promotes interactions with the factor XIII active site region. J
Thromb Haemost 7:1947–1949, 2009.
64. Cleary DB, Maurer MC. Characterizing the specificity of activated Factor XIII for
glutamine-containing substrate peptides. Biochim Biophys Acta 1764:1207–1217,
2006.
65. Cohen I, Blankenberg TA, Borden D, Kahn DR, Veis A. Factor XIIIa-catalyzed crosslinking of platelet and muscle actin. Regulation by nucleotides. Biochim Biophys Acta
628:365–375, 1980.
78. Credo RB, Curtis CG, Lorand L. Alpha-chain domain of fibrinogen controls generation
of fibrinoligase (coagulation factor XIIIa). Calcium ion regulatory aspects. Biochemistry
20:3770 –3778, 1981.
79. Credo RB, Curtis CG, Lorand L. Ca2⫹-related regulatory function of fibrinogen. Proc
Natl Acad Sci USA 75:4234 – 4237, 1978.
80. Csosz E, Mesko B, Fesus L. Transdab wiki: the interactive transglutaminase substrate
database on web 2.0 surface. Amino Acids 36:615– 617, 2009.
81. Curtis CG, Brown KL, Credo RB, Domanik RA, Gray A, Stenberg P, Lorand L.
Calcium-dependent unmasking of active center cysteine during activation of fibrin
stabilizing factor. Biochemistry 13 3774 –3780, 1974.
82. Czabanka M, Martin E, Walther A. Role of antithrombin and factor XIII in leukocyteindependent plasma extravasation during endotoxemia: an intravital-microscopic
study in the rat. J Surg Res 136:219 –226, 2006.
83. D’Argenio G, Grossman A, Cosenza V, Valle ND, Mazzacca G, Bishop PD. Recombinant factor XIII improves established experimental colitis in rats. Dig Dis Sci 45:987–
997, 2000.
84. Dallabrida SM, Falls LA, Farrell DH. Factor XIIIa supports microvascular endothelial
cell adhesion and inhibits capillary tube formation in fibrin. Blood 95:2586 –2592, 2000.
85. Dardik R, Krapp T, Rosenthal E, Loscalzo J, Inbal A. Effect of FXIII on monocyte and
fibroblast function. Cell Physiol Biochem 19:113–120, 2007.
86. Dardik R, Leor J, Skutelsky E, Castel D, Holbova R, Schiby G, Shaish A, Dickneite
G, Loscalzo J, Inbal A. Evaluation of the pro-angiogenic effect of factor XIII in
heterotopic mouse heart allografts and FXIII-deficient mice. Thromb Haemost
95:546 –550, 2006.
66. Cohen I, Gerrard JM, White JG. Ultrastructure of clots during isometric contraction.
J Cell Biol 93:775–787, 1982.
87. Dardik R, Livnat T, Seligsohn U. Variable effects of alpha v suppression on VEGFR-2
expression in endothelial cells of different vascular beds. Thromb Haemost 102:975–
982, 2009.
67. Cohen I, Glaser T, Veis A, Bruner-Lorand J. Ca2⫹-dependent cross-linking processes
in human platelets. Biochim Biophys Acta 676:137–147, 1981.
88. Dardik R, Loscalzo J, Eskaraev R, Inbal A. Molecular mechanisms underlying the
proangiogenic effect of factor XIII. Arterioscler Thromb Vasc Biol 25:526 –532, 2005.
68. Cohen I, Lim CT, Kahn DR, Glaser T, Gerrard JM, White JG. Disulfide-linked and
transglutaminase-catalyzed protein assemblies in platelets. Blood 66:143–151, 1985.
69. Cohen I, Young-Bandala L, Blankenberg TA, Siefring GE ,Jr, Bruner-Lorand J. Fibrinoligase-catalyzed cross-linking of myosin from platelet and skeletal muscle. Arch
Biochem Biophys 192:100 –111, 1979.
70. Colman RW, Clowes AW, George JN, Goldhaber SZ, Marder VJ. Overview of hemostasis. In: Hemostasis and Thrombosis, edited by Colman RW, Marder VJ, Clowes AW,
George JN, Goldhaber SZ. Philadelphia, PA:Lippincott Williams and Wilkins, 2006, p.
3–100 –20.
71. Colman RW, Clowes AW, George JN, Goldhaber SZ, Marder VJ. Thrombin-activatable fibrinolysis inhibitor aka procarboxypeptidase U. In: Hemostasis and Thrombosis,
edited by Colman RW, Marder VJ, Clowes AW, George JN, Goldhaber SZ. Philadelphia, PA:Lippincott Williams and Wilkins, 2006, p. 381–100 – 407.
72. Conkling PR, Achyuthan KE, Greenberg CS, Newcomb TF, Weinberg JB. Human
mononuclear phagocyte transglutaminase activity cross-links fibrin. Thromb Res 55:
57– 68, 1989.
73. Coopland A, Alkjaersig N, Fletcher AP. Reduction in plasma factor 13 (fibrin stabilizing
factor) concentration during pregnancy. J Lab Clin Med 73:144 –153, 1969.
74. Corbett SA, Lee L, Wilson CL, Schwarzbauer JE. Covalent cross-linking of fibronectin
to fibrin is required for maximal cell adhesion to a fibronectin-fibrin matrix. J Biol Chem
272:24999 –25005, 1997.
89. Dardik R, Shenkman B, Tamarin I, Eskaraev R, Harsfalvi J, Varon D, Inbal A. Factor XIII
mediates adhesion of platelets to endothelial cells through alpha(v)beta(3) and glycoprotein IIb/IIIa integrins. Thromb Res 105:317–323, 2002.
90. Dardik R, Solomon A, Loscalzo J, Eskaraev R, Bialik A, Goldberg I, Schiby G, Inbal A.
Novel proangiogenic effect of factor XIII associated with suppression of thrombospondin 1 expression. Arterioscler Thromb Vasc Biol 23:1472–1477, 2003.
91. De Willige SU, Standeven KF, Philippou H, Ariens RA. The pleiotropic role of the
fibrinogen gamma= chain in hemostasis. Blood 114:3994 – 4001, 2009.
92. Donga A, Kendrickb B, Kreilga˚rdb L, Matsuurab J, Manningb MC, Carpenterb JF.
Spectroscopic study of secondary structure and thermal denaturation of recombinant
human factor XIII in aqueous solution. Arch Biochem Biophys 347:213–220, 1997.
93. Doolittle RF. Searching for differences between fibrinogen and fibrin that affect the
initiation of fibrinolysis. Cardiovasc Hematol Agents Med Chem 6:181–189, 2008.
94. Duckert F, Jung E, Shmerling DH. A hitherto undescribed congenital haemorrhagic
diathesis probably due to fibrin stabilizing factor deficiency. Thromb Diath Haemorrh
5:179 –186, 1960.
95. Facchiano A, Facchiano F. Transglutaminases and their substrates in biology and
human diseases: 50 years of growing. Amino Acids 36:599 – 614, 2009.
96. Fairweather D, Cihakova D. Alternatively activated macrophages in infection and
autoimmunity. J Autoimmun 33:222–230, 2009.
75. Cordell PA, Kile BT, Standeven KF, Josefsson EC, Pease RJ, Grant PJ. Association of
coagulation factor XIII-A with Golgi proteins within monocyte-macrophages: implications for subcellular trafficking and secretion. Blood 115:2674 –2681, 2010.
97. Falls LA, Farrell DH. Resistance of gammaA/gamma= fibrin clots to fibrinolysis. J Biol
Chem 272:14251–14256, 1997.
76. Cornwell MM, Juliano RL, Davies PJ. Inhibition of the adhesion of Chinese hamster
ovary cells by the naphthylsulfonamides dansylcadaverine and N-(6-aminohexyl)-5chloro-1-naphthylenesulfonamide (W7). Biochim Biophys Acta 762:414 – 419, 1983.
98. Feng Q, Liu Y, Liu K, Byrne S, Liu G, Wang X, Li Z, Ockleford CD. Expression of
urokinase, plasminogen activator inhibitors and urokinase receptor in pregnant rhesus
monkey uterus during early placentation. Placenta 21:184 –193, 2000.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
965
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
61. Chung SI. Comparative studies on tissue transglutaminase and factor XIII. Ann NY Acad
Sci 202:240 –255, 1972.
77. Cottrell BA, Strong DD, Watt KW, Doolittle RF. Amino acid sequence studies on the
alpha chain of human fibrinogen. Exact location of cross-linking acceptor sites. Biochemistry 18:5405–5410, 1979.
MUSZBEK ET AL.
99. Fesus L, Metsis ML, Muszbek L, Koteliansky VE. Transglutaminase-sensitive glutamine
residues of human plasma fibronectin revealed by studying its proteolytic fragments.
Eur J Biochem 154:371–374, 1986.
121. Greenberg CS, Achyuthan KE, Fenton JW ,2nd. Factor XIIIa formation promoted
by complexing of alpha-thrombin, fibrin, and plasma factor XIII. Blood 69:867–
871, 1987.
100. Fisher S, Rikover M, Naor S. Factor 13 deficiency with severe hemorrhagic diathesis.
Blood 28:34 –39, 1966.
122. Greenberg CS, Achyuthan KE, Rajagopalan S, Pizzo SV. Characterization of the fibrin
polymer structure that accelerates thrombin cleavage of plasma factor XIII. Arch
Biochem Biophys 262:142–148, 1988.
101. Floridon C, Nielsen O, Holund B, Sunde L, Westergaard JG, Thomsen SG, Teisner B.
Localization and significance of urokinase plasminogen activator and its receptor in placental tissue from intrauterine, ectopic and molar pregnancies. Placenta 20:711–721,
1999.
102. Fox BA, Yee VC, Pedersen LC, Le Trong I, Bishop PD, Stenkamp RE, Teller DC.
Identification of the calcium binding site and a novel ytterbium site in blood coagulation factor XIII by x-ray crystallography. J Biol Chem 274:4917– 4923, 1999.
103. Francis CW, Marder VJ. Increased resistance to plasmic degradation of fibrin with
highly crosslinked alpha-polymer chains formed at high factor XIII concentrations.
Blood 71:1361–1365, 1988.
105. Francis RT, McDonagh J, Mann KG. Factor V is a substrate for the transamidase factor
XIIIa. J Biol Chem 261:9787–9792, 1986.
106. Gaffney PJ, Whitaker AN. Fibrin crosslinks and lysis rates. Thromb Res 14:85–94, 1979.
107. Gallivan L, Markham AF, Anwar R. The Leu564 factor XIIIA variant results in significantly lower plasma factor XIII levels than the Pro564 variant. Thromb Haemost 82:
1368 –1370, 1999.
108. Gao J, Hooker BS, Anderson DB. Expression of functional human coagulation factor
XIII A-domain in plant cell suspensions and whole plants. Protein Expr Purif 37:89 –96,
2004.
109. Gascuel O. BIONJ: an improved version of the NJ algorithm based on a simple model
of sequence data. Mol Biol Evol 14:685– 695, 1997.
110. Gemmati D, Federici F, Campo G, Tognazzo S, Serino ML, De Mattei M, Valgimigli M,
Malagutti P, Guardigli G, Ferraresi P, Bernardi F, Ferrari R, Scapoli GL, Catozzi L.
Factor XIIIA-V34L and factor XIIIB-H95R gene variants: effects on survival in myocardial infarction patients. Mol Med 13:112–120, 2007.
111. Gemmati D, Federici F, Catozzi L, Gianesini S, Tacconi G, Scapoli GL, Zamboni P.
DNA-array of gene variants in venous leg ulcers: detection of prognostic indicators. J
Vasc Surg 50:1444 –1451, 2009.
112. Gemmati D, Tognazzo S, Catozzi L, Federici F, De Palma M, Gianesini S, Scapoli GL,
De Mattei M, Liboni A, Zamboni P. Influence of gene polymorphisms in ulcer healing
process after superficial venous surgery. J Vasc Surg 44:554 –562, 2006.
113. Gemmati D, Tognazzo S, Serino ML, Fogato L, Carandina S, De Palma M, Izzo M, De
Mattei M, Ongaro A, Scapoli GL, Caruso A, Liboni A, Zamboni P. Factor XIII V34L
polymorphism modulates the risk of chronic venous leg ulcer progression and extension. Wound Repair Regen 12:512–517, 2004.
114. Gierhake FW, Papastavrou N, Zimmermann K, Bohn H, Schwick HG. Prophylaxis of
post-operative disturbances of wound healing with factor XIII substitution (author’s
transl). Dtsch Med Wochenschr 99:1004 –1009, 1974.
115. Gierhake FW, Volkmann W, Becker W, Schwarz H, Schwick HG. Factor XIII concentration and wound healing. Dtsch Med Wochenschr 95:1472–1475, 1970.
116. Girolami A, Cappellato MG, Lazzaro AR, Boscaro M. Type I and type II disease in
congenital factor XIII deficiency. A further demonstration of the correctness of the
classification. Blut 53:411– 413, 1986.
124. Greenberg CS, Sane DC, Lai TS. Factor XIII and fibrin stabilization. In: Hemostasis and
Thrombosis, edited by Colman RW, Clowes AW, Goldhaber SZ, Marder VJ, George
JN. Philadelphia, PA:Lippincott Williams and Wilkins, 2006, p. 317–466 –334.
125. Greenberg CS, Shuman MA. The zymogen forms of blood coagulation factor XIII bind
specifically to fibrinogen. J Biol Chem 257:6096 – 6101, 1982.
126. Grinnell F, Feld M, Minter D. Fibroblast adhesion to fibrinogen and fibrin substrata:
requirement for cold-insoluble globulin (plasma fibronectin). Cell 19:517–525, 1980.
127. Grundmann U, Amann E, Zettlmeissl G, Kupper HA. Characterization of cDNA
coding for human factor XIIIa. Proc Natl Acad Sci USA 83:8024 – 8028, 1986.
128. Harsfalvi J, Fesus L, Tarcsa E, Laczko J, Loewy AG. The presence of a covalently
cross-linked matrix in human platelets. Biochim Biophys Acta 1073:268 –274, 1991.
129. Hayano Y, Imai N, Karasawa T. Studies on the physiological changes of blood coagulation factor XIII during pregnancy and their significance. Nippon Sanka Fujinka Gakkai
Zasshi 34:469 – 477, 1982.
130. Henriksson P, Becker S, Lynch G, McDonagh J. Identification of intracellular factor XIII
in human monocytes and macrophages. J Clin Invest 76:528 –534, 1985.
131. Herouy Y, Hellstern MO, Vanscheidt W, Schopf E, Norgauer J. Factor XIII-mediated
inhibition of fibrinolysis and venous leg ulcers. Lancet 355:1970 –1971, 2000.
132. Hevessy Z, Haramura G, Boda Z, Udvardy M, Muszbek L. Promotion of the crosslinking of fibrin and alpha 2-antiplasmin by platelets. Thromb Haemost 75:161–167,
1996.
133. Higaki S, Nakano K, Onaka S, Amano A, Tanioka Y, Harada K, Hashimoto S, Sakaida
I, Okita K. Clinical significance of measuring blood coagulation factor XIIIA regularly
and continuously in patients with Crohn’s disease. J Gastroenterol Hepatol 21:1407–
1411, 2006.
134. Hildenbrand T, Idzko M, Panther E, Norgauer J, Herouy Y. Treatment of nonhealing
leg ulcers with fibrin-stabilizing factor XIII: a case report. Dermatol Surg 28:1098 –
1099, 2002.
135. Hirahara K, Shinbo K, Takahashi M, Matsuishi T. Suppressive effect of human blood
coagulation factor XIII on the vascular permeability induced by anti-guinea pig endothelial cell antiserum in guinea pigs. Thromb Res 71:139 –148, 1993.
136. Holmes WE, Lijnen HR, Collen D. Characterization of recombinant human alpha
2-antiplasmin and of mutants obtained by site-directed mutagenesis of the reactive
site. Biochemistry 26:5133–5140, 1987.
137. Hornyak TJ, Shafer JA. Interactions of factor XIII with fibrin as substrate and cofactor.
Biochemistry 31:423– 429, 1992.
138. Hornyak TJ, Shafer JA. Role of calcium ion in the generation of factor XIII activity.
Biochemistry 30:6175– 6182, 1991.
139. Hsieh L, Nugent D. Factor XIII deficiency. Haemophilia 14:1190 –1200, 2008.
117. Gladner JA, Nossal R. Effects of crosslinking on the rigidity and proteolytic susceptibility of human fibrin clots. Thromb Res 30:273–288, 1983.
140. Huh MM, Schick BP, Schick PK, Colman RW. Covalent crosslinking of human coagulation factor V by activated factor XIII from guinea pig megakaryocytes and human
plasma. Blood 71:1693–1702, 1988.
118. Glaros T, Larsen M, Li L. Macrophages and fibroblasts during inflammation, tissue
damage and organ injury. Front Biosci 14:3988 –3993, 2009.
141. Huson DH, Bryant D. Application of phylogenetic networks in evolutionary studies.
Mol Biol Evol 23:254 –267, 2006.
119. Glynn LE (Editor) . Tissue Repair and Regeneration. New York:Raven, 1985.
142. Ichinose A, Asahina T, Kobayashi T. Congenital blood coagulation factor XIII deficiency and perinatal management. Curr Drug Targets 6:541–549, 2005.
120. Gratchev A, Kzhyshkowska J, Utikal J, Goerdt S. Interleukin-4 and dexamethasone
counterregulate extracellular matrix remodelling and phagocytosis in type-2 macrophages. Scand J Immunol 61:10 –17, 2005.
966
143. Ichinose A, Bottenus RE, Davie EW. Structure of transglutaminases. J Biol Chem 265:
13411–13414, 1990.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
104. Francis CW, Marder VJ. Rapid formation of large molecular weight alpha-polymers in
cross-linked fibrin induced by high factor XIII concentrations. Role of platelet factor
XIII. J Clin Invest 80:1459 –1465, 1987.
123. Greenberg CS, Miraglia CC. The effect of fibrin polymers on thrombin-catalyzed
plasma factor XIIIa formation. Blood 66:466 – 469, 1985.
MULTIPLE FUNCTIONS OF FACTOR XIII
144. Ichinose A, Davie EW. Characterization of the gene for the a subunit of human factor
XIII (plasma transglutaminase), a blood coagulation factor. Proc Natl Acad Sci USA
85:5829 –5833, 1988.
165. Johnson K, Hashimoto S, Lotz M, Pritzker K, Terkeltaub R. Interleukin-1 induces
pro-mineralizing activity of cartilage tissue transglutaminase and factor XIIIa. Am J
Pathol 159:149 –163, 2001.
145. Ichinose A, Hendrickson LE, Fujikawa K, Davie EW. Amino acid sequence of the a
subunit of human factor XIII. Biochemistry 25:6900 – 6906, 1986.
166. Johnson KA, Rose DM, Terkeltaub RA. Factor XIIIA mobilizes transglutaminase 2 to
induce chondrocyte hypertrophic differentiation. J Cell Sci 121:2256 –2264, 2008.
146. Ichinose A, McMullen BA, Fujikawa K, Davie EW. Amino acid sequence of the b
subunit of human factor XIII, a protein composed of ten repetitive segments. Biochemistry 25:4633– 4638, 1986.
167. Joist JH, Niewiarowski S. Retention of platelet fibrin stabilizing factor during the
platelet release reaction and clot retraction. Thromb Diath Haemorrh 29:679 – 683,
1973.
147. Iismaa SE, Holman S, Wouters MA, Lorand L, Graham RM, Husain A. Evolutionary
specialization of a tryptophan indole group for transition-state stabilization by eukaryotic transglutaminases. Proc Natl Acad Sci USA 100:12636 –12641, 2003.
168. Kaartinen MT, El-Maadawy S, Rasanen NH, McKee MD. Tissue transglutaminase and
its substrates in bone. J Bone Miner Res 17:2161–2173, 2002.
148. Iismaa SE, Mearns BM, Lorand L, Graham RM. Transglutaminases and disease: lessons
from genetically engineered mouse models and inherited disorders. Physiol Rev 89:
991–1023, 2009.
169. Kaetsu H, Hashiguchi T, Foster D, Ichinose A. Expression and release of the a and b
subunits for human coagulation factor XIII in baby hamster kidney (BHK) cells. J
Biochem 119:961–969, 1996.
170. Kappelmayer J, Bacsko G, Birinyi L, Zakany R, Kelemen E, Adany R. Consecutive
appearance of coagulation factor XIII subunit A in macrophages, megakaryocytes,
liver cells during early human development. Blood 86:2191–2197, 1995.
150. Inbal A, Lubetsky A, Krapp T, Castel D, Shaish A, Dickneitte G, Modis L, Muszbek L.
Impaired wound healing in factor XIII deficient mice. Thromb Haemost 94:432– 437,
2005.
171. Kappelmayer J, Bacsko G, Kelemen E, Adany R. Onset and distribution of factor
XIII-containing cells in the mesenchyme of chorionic villi during early phase of human
placentation. Placenta 15:613– 623, 1994.
151. Inbal A, Muszbek L. Coagulation factor deficiencies and pregnancy loss. Semin Thromb
Hemost 29:171–174, 2003.
172. Kappelmayer J, Simon A, Katona E, Szanto A, Nagy L, Kiss A, Kiss C, Muszbek L.
Coagulation factor XIII-A. A flow cytometric intracellular marker in the classification
of acute myeloid leukemias. Thromb Haemost 94:454 – 459, 2005.
152. Inbal A, Muszbek L, Lubetsky A, Katona E, Levi I, Karpati L, Nagler A. Platelets but not
monocytes contribute to the plasma levels of factor XIII subunit A in patients undergoing autologous peripheral blood stem cell transplantation. Blood Coagul Fibrinolysis
15:249 –253, 2004.
173. Karimi M, Bereczky Z, Cohan N, Muszbek L. Factor XIII deficiency. Semin Thromb
Hemost 35:426 – 438, 2009.
153. Isogai N, Ueda Y, Kurozumi N, Kamiishi H. Wound healing at the site of microvascular
anastomosis: fibrin-stabilizing factor XIII administration and its effects. Microsurgery
11:40 – 46, 1990.
154. Ivaskevicius V, Seitz R, Kohler HP, Schroeder V, Muszbek L, Ariens RA, Seifried E,
Oldenburg J. International registry on factor XIII deficiency: a basis formed mostly on
European data. Thromb Haemost 97:914 –921, 2007.
155. Iwata H, Kitano T, Umetsu K, Yuasa I, Yamazaki K, Kemkes-Matthes B, Ichinose A.
Distinct C-terminus of the B subunit of factor XIII in a population-associated major
phenotype: the first case of complete allele-specific alternative splicing products in the
coagulation and fibrinolytic systems. J Thromb Haemost 7:1084 –1091, 2009.
156. Jaffer FA, Sosnovik DE, Nahrendorf M, Weissleder R. Molecular imaging of myocardial
infarction. J Mol Cell Cardiol 41:921–933, 2006.
157. Jagadeeswaran P, Reddy SV, Haas P. Synthesis of human coagulation factor XIII in
yeast. Gene 86:279 –283, 1990.
158. Jansen JW, Haverkate F, Koopman J, Nieuwenhuis HK, Kluft C, Boschman TA. Influence of factor XIIIa activity on human whole blood clot lysis in vitro. Thromb Haemost
57:171–175, 1987.
159. Janus TJ, Lewis SD, Lorand L, Shafer JA. Promotion of thrombin-catalyzed activation
of factor XIII by fibrinogen. Biochemistry 22:6269 – 6272, 1983.
174. Kasahara K, Souri M, Kaneda M, Miki T, Yamamoto N, Ichinose A. Impaired clot
retraction in factor XIII A subunit-deficient mice. Blood 115:1277–1279, 2010.
175. Katona E, Haramura G, Karpati L, Fachet J, Muszbek L. A simple, quick one-step
ELISA assay for the determination of complex plasma factor XIII (A2B2). Thromb
Haemost 83:268 –273, 2000.
176. Katona EE, Ajzner E, Toth K, Karpati L, Muszbek L. Enzyme-linked immunosorbent
assay for the determination of blood coagulation factor XIII A-subunit in plasma and in
cell lysates. J Immunol Methods 258:127–135, 2001.
177. Kavai M, Adany R, Pasti G, Suranyi P, Szucs G, Muszbek L, Bojan F, Szegedi G. Marker
profile, enzyme activity, and function of a human myelomonocytic leukemia cell line.
Cell Immunol 139:531–540, 1992.
178. Kida M, Souri M, Yamamoto M, Saito H, Ichinose A. Transcriptional regulation of cell
type-specific expression of the TATA-less A subunit gene for human coagulation
factor XIII. J Biol Chem 274:6138 – 6147, 1999.
179. Kiesselbach TH, Wagner RH. Demonstration of factor XIII in human megakaryocytes
by a fluorescent antibody technique. Ann NY Acad Sci 202:318 –328, 1972.
180. Kilian O, Fuhrmann R, Alt V, Noll T, Coskun S, Dingeldein E, Schnettler R, Franke RP.
Plasma transglutaminase factor XIII induces microvessel ingrowth into biodegradable
hydroxyapatite implants in rats. Biomaterials 26:1819 –1827, 2005.
181. Kimura S, Aoki N. Cross-linking site in fibrinogen for alpha 2-plasmin inhibitor. J Biol
Chem 261:15591–15595, 1986.
160. Jayo A, Conde I, Lastres P, Jimenez-Yuste V, Gonzalez-Manchon C. New insights into
the expression and role of platelet factor XIII-A. J Thromb Haemost 7:1184 –1191,
2009.
182. Kimura S, Tamaki T, Aoki N. Acceleration of fibrinolysis by the N-terminal peptide of
alpha 2-plasmin inhibitor. Blood 66:157–160, 1985.
161. Jayo A, Conde I, Lastres P, Jimenez-Yuste V, Gonzalez-Manchon C. Possible role for
cellular FXIII in monocyte-derived dendritic cell motility. Eur J Cell Biol 88:423– 431,
2009.
183. Kiss F, Hevessy Z, Veszpremi A, Katona E, Kiss C, Vereb G, Muszbek L, Kappelmayer
JN. Leukemic lymphoblasts, a novel expression site of coagulation factor XIII subunit
A. Thromb Haemost 96:176 –182, 2006.
162. Jelenska M, Kopec M, Breddin K. On the retraction of collagen and fibrin induced by
normal, defective and modified platelets. Haemostasis 15:169 –175, 1985.
184. Kobayashi T, Asahina T, Okada Y, Terao T. Studies on the localization of adhesive
proteins associated with the development of extravillous cytotrophoblast. Trophoblast
Res 13:35–53, 1999.
163. Jensen PH, Lorand L, Ebbesen P, Gliemann J. Type-2 plasminogen-activator inhibitor
is a substrate for trophoblast transglutaminase and factor XIIIa. Transglutaminasecatalyzed cross-linking to cellular and extracellular structures. Eur J Biochem 214:141–
146, 1993.
185. Kobayashi T, Terao T, Kojima T, Takamatsu J, Kamiya T, Saito H. Congenital factor
XIII deficiency with treatment of factor XIII concentrate and normal vaginal delivery.
Gynecol Obstet Invest 29:235–238, 1990.
164. Jobe SM, Leo L, Eastvold JS, Dickneite G, Ratliff TL, Lentz SR, Di Paola J. Role of
FcRgamma and factor XIIIA in coated platelet formation. Blood 106:4146 – 4151,
2005.
186. Komanasin N, Catto AJ, Futers TS, van Hylckama Vlieg A, Rosendaal FR, Ariens RA. A
novel polymorphism in the factor XIII B-subunit (His95Arg): relationship to subunit
dissociation and venous thrombosis. J Thromb Haemost 3:2487–2496, 2005.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
967
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
149. Ikematsu S. An approach to the metabolism of factor XIII. Nippon Ketsueki Gakkai
Zasshi 44:1499 –1505, 1981.
MUSZBEK ET AL.
209. Lewis BA, Freyssinet JM, Holbrook JJ. An equilibrium study of metal ion binding to
human plasma coagulation factor XIII. Biochem J 169:397– 402, 1978.
188. Kopec M, Latallo ZS, Stahl M, Wegrzynowicz Z. The effect of proteolytic enzymes on
fibrin stabilizing factor. Biochim Biophys Acta 181:437– 445, 1969.
210. Lewis SD, Janus TJ, Lorand L, Shafer JA. Regulation of formation of factor XIIIa by its
fibrin substrates. Biochemistry 24:6772– 6777, 1985.
189. Koseki-Kuno S, Yamakawa M, Dickneite G, Ichinose A. Factor XIII A subunit-deficient
mice developed severe uterine bleeding events and subsequent spontaneous miscarriages. Blood 102:4410 – 4412, 2003.
211. Lim BC, Ariens RA, Carter AM, Weisel JW, Grant PJ. Genetic regulation of fibrin
structure and function: complex gene-environment interactions may modulate vascular risk. Lancet 361:1424 –1431, 2003.
190. Kradin RL, Lynch GW, Kurnick JT, Erikson M, Colvin RB, McDonagh J. Factor XIII A is
synthesized and expressed on the surface of U937 cells and alveolar macrophages.
Blood 69:778 –785, 1987.
212. Liu JN, Kung W, Harpel PC, Gurewich V. Demonstration of covalent binding of
lipoprotein(a) [Lp(a)] to fibrin and endothelial cells. Biochemistry 37:3949 –3954,
1998.
191. Krarup A, Gulla KC, Gal P, Hajela K, Sim RB. The action of MBL-associated serine
protease 1 (MASP1) on factor XIII and fibrinogen. Biochim Biophys Acta 1784:1294 –
1300, 2008.
213. Liu S, Cerione RA, Clardy J. Structural basis for the guanine nucleotide-binding activity
of tissue transglutaminase and its regulation of transamidation activity. Proc Natl Acad
Sci USA 99:2743–2747, 2002.
192. Kulkarni S, Jackson SP. Platelet factor XIII and calpain negatively regulate integrin
alphaIIbbeta3 adhesive function and thrombus growth. J Biol Chem 279:30697–30706,
2004.
214. Lockwood CR, Frayling TM. Combining genome and mouse knockout expression
data to highlight binding sites for the transcription factor HNF1alpha. In Silico Biol
3:57–70, 2003.
193. Kurochkin IV, Procyk R, Bishop PD, Yee VC, Teller DC, Ingham KC, Medved LV.
Domain structure, stability and domain-domain interactions in recombinant factor
XIII. J Mol Biol 248:414 – 430, 1995.
215. Loewy AG, Dahlberg A, Dunathan K, Kriel R, Wolfinger HL ,Jr. Fibrinase. II. Some
physical properties. J Biol Chem 236:2634 –2643, 1961.
194. Lahav J, Karniel E, Bagoly Z, Sheptovitsky V, Dardik R, Inbal A. Coagulation factor XIII
serves as protein disulfide isomerase. Thromb Haemost 101:840 – 844, 2009.
195. Lai TS, Slaughter TF, Peoples KA, Greenberg CS. Site-directed mutagenesis of the
calcium-binding site of blood coagulation factor XIIIa. J Biol Chem 274:24953–24958,
1999.
216. Loewy AG, Dunathan K, Gallant JA, Gardner B. Fibrinase. III. Some enzymatic properties. J Biol Chem 236:2644 –2647, 1961.
217. Loewy AG, Dunathan K, Kriel R, Wolfinger HL ,Jr. Fibrinase. I. Purification of substrate
and enzyme. J Biol Chem 236:2625–2633, 1961.
218. Loewy AG, Gallant JA, Dunathan K. Fibrinase. IV. Effect on fibrin solubility. J Biol Chem
236:2648 –2655, 1961.
196. Laki K, Lorand L. On the solubility of fibrin clots. Science 108:280, 1948.
197. Lanir N, Ciano PS, Van de Water L, McDonagh J, Dvorak AM, Dvorak HF. Macrophage migration in fibrin gel matrices. II. Effects of clotting factor XIII, fibronectin,
glycosaminoglycan content on cell migration. J Immunol 140:2340 –2349, 1988.
198. Larkin MA, Blackshields G, Brown NP, Chenna R, McGettigan PA, McWilliam H,
Valentin F, Wallace IM, Wilm A, Lopez R, Thompson JD, Gibson TJ, Higgins DG.
Clustal W and Clustal X version 2.0. Bioinformatics 23:2947–2948, 2007.
219. Loewy AG, Veneziale C, Forman M. Purification of the factor involved in the formation of urea-insoluble fibrin. Biochim Biophys Acta 26:670 – 671, 1957.
220. Lopaciuk S, Lovette KM, McDonagh J, Chuang HY. Subcellular distribution of fibrinogen and factor XIII in human blood platelets. Thromb Res 8:453– 465, 1976.
221. Lorand L. A study on the solubility of fibrin clots in urea. Hung Acta Physiol 1:192–196,
1948.
199. Lauer P, Metzner HJ, Zettlmeissl G, Li M, Smith AG, Lathe R, Dickneite G. Targeted
inactivation of the mouse locus encoding coagulation factor XIII-A: hemostatic abnormalities in mutant mice and characterization of the coagulation deficit. Thromb Haemost 88:967–974, 2002.
222. Lorand L. Fibrin clots. Nature 166:694 – 695, 1950.
200. Lawler J. Thrombospondin-1 as an endogenous inhibitor of angiogenesis and tumor
growth. J Cell Mol Med 6:1–12, 2002.
224. Lorand L, Jeong JM, Radek JT, Wilson J. Human plasma factor XIII: subunit interactions
and activation of zymogen. Methods Enzymol 222:22–35, 1993.
201. Lee IH, Chung SI, Lee SY. Effects of Val34Leu and Val35Leu polymorphism on the
enzyme activity of the coagulation factor XIII-A. Exp Mol Med 34:385–390, 2002.
225. Lorand L, Konishi K, Jacobsen A. Transpeptidation mechanism in blood clotting.
Nature 194:1148 –1149, 1962.
202. Lee KN, Jackson KW, Christiansen VJ, Chung KH, McKee PA. A novel plasma proteinase potentiates alpha2-antiplasmin inhibition of fibrin digestion. Blood 103:3783–
3788, 2004.
226. Lorenz R, Born P, Olbert P, Classen M. Factor XIII substitution in ulcerative colitis.
Lancet 345:449 – 450, 1995.
203. Lee KN, Jackson KW, Christiansen VJ, Lee CS, Chun JG, McKee PA. Why alphaantiplasmin must be converted to a derivative form for optimal function. J Thromb
Haemost 5:2095–2104, 2007.
204. Lee KN, Lee CS, Tae WC, Jackson KW, Christiansen VJ, McKee PA. Cross-linking of
wild-type and mutant alpha 2-antiplasmins to fibrin by activated factor XIII and by a
tissue transglutaminase. J Biol Chem 275:37382–37389, 2000.
205. Lee KN, Lee SC, Jackson KW, Tae WC, Schwartzott DG, McKee PA. Effect of
phenylglyoxal-modified alpha2-antiplasmin on urokinase-induced fibrinolysis. Thromb
Haemost 80:637– 644, 1998.
206. Lee KN, Tae WC, Jackson KW, Kwon SH, McKee PA. Characterization of wild-type
and mutant alpha2-antiplasmins: fibrinolysis enhancement by reactive site mutant.
Blood 94:164 –171, 1999.
223. Lorand L, Credo RB, Janus TJ. Factor XIII (fibrin-stabilizing factor). Methods Enzymol
80:333–341, 1981.
227. Loscalzo J, Weinfeld M, Fless GM, Scanu AM. Lipoprotein(a), fibrin binding, plasminogen activation. Arteriosclerosis 10:240 –245, 1990.
228. Luscher EF. Fibrin-stabilizing factor from thrombocytes. Schweiz Med Wochenschr
87:1220 –1221, 1957.
229. Lynch GW, Slayter HS, Miller BE, McDonagh J. Characterization of thrombospondin
as a substrate for factor XIII transglutaminase. J Biol Chem 262:1772–1778, 1987.
230. Ma J, Liew CC. Gene profiling identifies secreted protein transcripts from peripheral
blood cells in coronary artery disease. J Mol Cell Cardiol 35:993–998, 2003.
231. Malara A, Gruppi C, Rebuzzini P, Visai L, Perotti C, Moratti R, Balduini C, Tira ME,
Balduini A. Megakaryocyte-matrix interaction within bone marrow: new roles for
fibronectin and factor XIII-A. Blood. In press.
207. Leifheit HJ, Cleve H. Analysis of the genetic polymorphism of coagulation factor XIIIB
(FXIIIB) by isoelectric focusing. Electrophoresis 9:426 – 429, 1988.
232. Marinescu A, Cleary DB, Littlefield TR, Maurer MC. Structural features associated
with the binding of glutamine-containing peptides to Factor XIII. Arch Biochem Biophys
406:9 –20, 2002.
208. Lesort M, Attanavanich K, Zhang J, Johnson GV. Distinct nuclear localization and
activity of tissue transglutaminase. J Biol Chem 273:11991–11994, 1998.
233. Marx G, Korner G, Mou X, Gorodetsky R. Packaging zinc, fibrinogen, factor XIII in
platelet alpha-granules. J Cell Physiol 156:437– 442, 1993.
968
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
187. Komaromi I, Bagoly Z, Muszbek L. Factor XIII: novel structural and functional aspects.
J Thromb Haemost 9:9 –20, 2011.
MULTIPLE FUNCTIONS OF FACTOR XIII
234. Matsuka YV, Medved LV, Migliorini MM, Ingham KC. Factor XIIIa-catalyzed crosslinking of recombinant alpha C fragments of human fibrinogen. Biochemistry 35:5810 –
5816, 1996.
255. Muszbek L, Adany R, Kavai M, Boda Z, Lopaciuk S. Monocytes of patients congenitally
deficient in plasma factor XIII lack factor XIII subunit a antigen and transglutaminase
activity. Thromb Haemost 59:231–235, 1988.
235. McDonagh J, Fukue H. Determinants of substrate specificity for factor XIII. Semin
Thromb Hemost 22:369 –376, 1996.
256. Muszbek L, Adany R, Mikkola H. Novel aspects of blood coagulation factor XIII. I.
Structure, distribution, activation, and function. Crit Rev Clin Lab Sci 33:357– 421,
1996.
236. McDonagh J, McDonagh RP. Alternative pathways for the activation of factor XIII. Br
J Haematol 30:465– 477, 1975.
237. McDonagh RP ,Jr, McDonagh J, Duckert F. The influence of fibrin crosslinking on the
kinetics of urokinase-induced clot lysis. Br J Haematol 21:323–332, 1971.
238. McDonagh RP, McDonagh J, Petersen TE, Thogersen HC, Skorstengaard K, SottrupJensen L, Magnusson S, Dell A, Morris HR. Amino acid sequence of the factor XIIIa
acceptor site in bovine plasma fibronectin. FEBS Lett 127:174 –178, 1981.
257. Muszbek L, Adany R, Szegedi G, Polgar J, Kavai M. Factor XIII of blood coagulation in
human monocytes. Thromb Res 37:401– 410, 1985.
258. Muszbek L, Ariens RA, Ichinose A. Factor XIII: recommended terms and abbreviations. J Thromb Haemost 5:181–183, 2007.
259. Muszbek L, Bagoly Z, Bereczky Z, Katona E. The involvement of blood coagulation
factor XIII in fibrinolysis and thrombosis. Cardiovasc Hematol Agents Med Chem 6:190 –
205, 2008.
260. Muszbek L, Bereczky Z, Bagoly Z, Shemirani AH, Katona E.Factor XIII and atherothrombotic diseases. Semin Thromb Hemost 36:18 –33.
240. Mendelsohn ME, Zhu Y, O’Neill S. The 29-kDa proteins phosphorylated in thrombinactivated human platelets are forms of the estrogen receptor-related 27-kDa heat
shock protein. Proc Natl Acad Sci USA 88:11212–11216, 1991.
261. Muszbek L, Haramura G, Polgar J. Transformation of cellular factor XIII into an active
zymogen transglutaminase in thrombin-stimulated platelets. Thromb Haemost 73:
702–705, 1995.
241. Merz D, Liu R, Johnson K, Terkeltaub R. IL-8/CXCL8 and growth-related oncogene
alpha/CXCL1 induce chondrocyte hypertrophic differentiation. J Immunol 171:4406 –
4415, 2003.
262. Muszbek L, Polgar J, Boda Z. Platelet factor XIII becomes active without the release
of activation peptide during platelet activation. Thromb Haemost 69:282–285, 1993.
242. Metha K, Turpin J, Lopez-Berestein G. Induction of tissue transglutaminase in human
peripheral blood monocytes by intracellular delivery of retinoids. J Leukoc Biol 41:341–
348, 1987.
243. Mikkola H, Muszbek L, Laiho E, Syrjala M, Hamalainen E, Haramura G, Salmi T,
Peltonen L, Palotie A. Molecular mechanism of a mild phenotype in coagulation factor
XIII (FXIII) deficiency: a splicing mutation permitting partial correct splicing of FXIII
A-subunit mRNA. Blood 89:1279 –1287, 1997.
263. Muszbek L, Yee VC, Hevessy Z. Blood coagulation factor XIII: structure and function.
Thromb Res 94:271–305, 1999.
264. Mutch NJ, Koikkalainen JS, Fraser SR, Duthie KM, Griffin M, Mitchell J, Watson HG,
Booth NA. Model thrombi formed under flow reveal the role of factor XIII-mediated
cross-linking in resistance to fibrinolysis. J Thromb Haemost 8:2017–2024, 2010.
265. Nagy B ,Jr, Simon Z, Bagoly Z, Muszbek L, Kappelmayer J. Binding of plasma factor
XIII to thrombin-receptor activated human platelets. Thromb Haemost 102:83– 89,
2009.
244. Mikkola H, Syrjala M, Rasi V, Vahtera E, Hamalainen E, Peltonen L, Palotie A. Deficiency in the A-subunit of coagulation factor XIII: two novel point mutations demonstrate different effects on transcript levels. Blood 84:517–525, 1994.
266. Nagy JA, Henriksson P, McDonagh J. Biosynthesis of factor XIII B subunit by human
hepatoma cell lines. Blood 68:1272–1279, 1986.
245. Mirochnik Y, Kwiatek A, Volpert OV. Thrombospondin and apoptosis: molecular
mechanisms and use for design of complementation treatments. Curr Drug Targets
9:851– 862, 2008.
267. Nahrendorf M, Aikawa E, Figueiredo JL, Stangenberg L, van den Borne SW, Blankesteijn WM, Sosnovik DE, Jaffer FA, Tung CH, Weissleder R. Transglutaminase activity in acute infarcts predicts healing outcome and left ventricular remodelling: implications for FXIII therapy and antithrombin use in myocardial infarction. Eur Heart J
29:445– 454, 2008.
246. Mishima Y, Nagao F, Ishibiki K, Matsuda M, Nakamura N. Factor XIII in the treatment
of postoperative refractory wound-healing disorders. Results of a controlled study.
Chirurg 55:803– 808, 1984.
247. Moaddel M, Falls LA, Farrell DH. The role of gamma A/gamma= fibrinogen in plasma
factor XIII activation. J Biol Chem 275:32135–32140, 2000.
248. Moaddel M, Farrell DH, Dougherty MA, Frieol MG. Interactions of human fibrinogens
with factor XIII: roles of calcium and the gamma peptide. Biochemistry 39:6698 – 6705,
2000.
249. Mortensen SB, Sottrup-Jensen L, Hansen HF, Rider D, Petersen TE, Magnusson S.
Sequence location of a putative transglutaminase crosslinking site in human alpha
2-macroglobulin. FEBS Lett 129:314 –317, 1981.
268. Nahrendorf M, Hu K, Frantz S, Jaffer FA, Tung CH, Hiller KH, Voll S, Nordbeck P,
Sosnovik D, Gattenlohner S, Novikov M, Dickneite G, Reed GL, Jakob P, Rosenzweig
A, Bauer WR, Weissleder R, Ertl G. Factor XIII deficiency causes cardiac rupture,
impairs wound healing, and aggravates cardiac remodeling in mice with myocardial
infarction. Circulation 113:1196 –1202, 2006.
269. Nahrendorf M, Sosnovik DE, Waterman P, Swirski FK, Pande AN, Aikawa E,
Figueiredo JL, Pittet MJ, Weissleder R. Dual channel optical tomographic imaging of
leukocyte recruitment and protease activity in the healing myocardial infarct. Circ Res
100:1218 –1225, 2007.
270. Nahrendorf M, Weissleder R, Ertl G. Does FXIII deficiency impair wound healing after
myocardial infarction? PLoS One 1:e48, 2006.
250. Mosesson MW. Fibrinogen and fibrin structure and functions. J Thromb Haemost
3:1894 –1904, 2005.
271. Nakano Y, Al-Jallad HF, Mousa A, Kaartinen MT. Expression and localization of plasma
transglutaminase factor XIIIA in bone. J Histochem Cytochem 55:675– 685, 2007.
251. Mosesson MW, Siebenlist KR, Amrani DL, DiOrio JP. Identification of covalently
linked trimeric and tetrameric D domains in crosslinked fibrin. Proc Natl Acad Sci USA
86:1113–1117, 1989.
272. Naski MC, Lorand L, Shafer JA. Characterization of the kinetic pathway for fibrin
promotion of alpha-thrombin-catalyzed activation of plasma factor XIII. Biochemistry
30:934 –941, 1991.
252. Mosesson MW, Siebenlist KR, Hernandez I, Lee KN, Christiansen VJ, McKee PA.
Evidence that alpha2-antiplasmin becomes covalently ligated to plasma fibrinogen in
the circulation: a new role for plasma factor XIII in fibrinolysis regulation. J Thromb
Haemost 6:1565–1570, 2008.
273. Niewiarowski S, Kirby EP, Brudzynski TM, Stocker K. Thrombocytin, a serine protease from Bothrops atrox venom. 2. Interaction with platelets and plasma-clotting
factors. Biochemistry 18:3570 –3577, 1979.
253. Mosnier LO, Bouma BN. Regulation of fibrinolysis by thrombin activatable fibrinolysis
inhibitor, an unstable carboxypeptidase B that unites the pathways of coagulation and
fibrinolysis. Arterioscler Thromb Vasc Biol 26:2445–2453, 2006.
254. Murtaugh MP, Arend WP, Davies PJ. Induction of tissue transglutaminase in human
peripheral blood monocytes. J Exp Med 159:114 –125, 1984.
274. Niewiarowski S, Markiewicz M, Nath N. Inhibition of the platelet-dependent fibrin
retraction by the fibrin stabilizing factor (FSF, factor 13). J Lab Clin Med 81:641– 650,
1973.
275. Noll T, Wozniak G, McCarson K, Hajimohammad A, Metzner HJ, Inserte J, Kummer
W, Hehrlein FW, Piper HM. Effect of factor XIII on endothelial barrier function. J Exp
Med 189:1373–1382, 1999.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
969
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
239. Mehta K, Lopez-Berestein G, Moore WT, Davies PJ. Interferon-gamma requires
serum retinoids to promote the expression of tissue transglutaminase in cultured
human blood monocytes. J Immunol 134:2053–2056, 1985.
MUSZBEK ET AL.
276. Nor JE, Mitra RS, Sutorik MM, Mooney DJ, Castle VP, Polverini PJ. Thrombospondin-1 induces endothelial cell apoptosis and inhibits angiogenesis by activating the
caspase death pathway. J Vasc Res 37:209 –218, 2000.
297. Peschen M, Thimm C, Weyl A, Weiss JM, Kurz H, Augustin M, Simon JC, Schopf E,
Vanscheidt W. Possible role of coagulation factor XIII in the pathogenesis of venous
leg ulcers. Vasa 27:89 –93, 1998.
277. Nossel HL, Lanzkowsky P, Levy S, Mibashan RS, Hansen JD. A study of coagulation
factor levels in women during labour and in their newborn infants. Thromb Diath
Haemorrh 16:185–197, 1966.
298. Peter A, Lilja H, Lundwall A, Malm J. Semenogelin I and semenogelin II, the major
gel-forming proteins in human semen, are substrates for transglutaminase. Eur J
Biochem 252:216 –221, 1998.
278. Nurden AT, Kunicki TJ, Dupuis D, Soria C, Caen JP. Specific protein and glycoprotein
deficiencies in platelets isolated from two patients with the gray platelet syndrome.
Blood 59:709 –718, 1982.
299. Pettersen EF, Goddard TD, Huang CC, Couch GS, Greenblatt DM, Meng EC, Ferrin
TE. UCSF chimera: a visualization system for exploratory research and analysis. J
Comput Chem 25:1605–1612, 2004.
279. Nurminskaya M, Kaartinen MT. Transglutaminases in mineralized tissues. Front Biosci
11:1591–1606, 2006.
300. Philippou H, Rance J, Myles T, Hall SW, Ariens RA, Grant PJ, Leung L, Lane DA. Roles
of low specificity and cofactor interaction sites on thrombin during factor XIII activation. Competition for cofactor sites on thrombin determines its fate. J Biol Chem
278:32020 –32026, 2003.
280. Nurminskaya M, Magee C, Nurminsky D, Linsenmayer TF. Plasma transglutaminase
in hypertrophic chondrocytes: expression and cell-specific intracellular activation produce cell death and externalization. J Cell Biol 142:1135–1144, 1998.
301. Pinkas DM, Strop P, Brunger AT, Khosla C. Transglutaminase 2 undergoes a large
conformational change upon activation. PLoS Biol 5:e327, 2007.
302. Polgar J, Hidasi V, Muszbek L. Non-proteolytic activation of cellular protransglutaminase (placenta macrophage factor XIII). Biochem J 267:557–560, 1990.
282. Nurminskaya MV, Recheis B, Nimpf J, Magee C, Linsenmayer TF. Transglutaminase
factor XIIIA in the cartilage of developing avian long bones. Dev Dyn 223:24 –32, 2002.
303. Poon MC, Russell JA, Low S, Sinclair GD, Jones AR, Blahey W, Ruether BA, Hoar DI.
Hemopoietic origin of factor XIII A subunits in platelets, monocytes, and plasma.
Evidence from bone marrow transplantation studies. J Clin Invest 84:787–792, 1989.
283. Ogasawara MS, Aoki K, Katano K, Ozaki Y, Suzumori K. Factor XII but not protein C,
protein S, antithrombin III, or factor XIII is a predictor of recurrent miscarriage. Fertil
Steril 75:916 –919, 2001.
304. Quatresooz P, Paquet P, Hermanns-Le T, Pierard GE. Molecular mapping of Factor
XIIIa-enriched dendrocytes in the skin (review). Int J Mol Med 22:403– 409, 2008.
284. Ogawa T, Morioka Y, Inoue T, Takano M, Tsuda S. Involvement of blood coagulation
factor XIII in burn healing in the carbon tetrachloride-induced hepatic injury model in
rats. Inflamm Res 44:264 –268, 1995.
285. Ohmori T, Yatomi Y, Asazuma N, Satoh K, Ozaki Y. Involvement of proline-rich
tyrosine kinase 2 in platelet activation: tyrosine phosphorylation mostly dependent on
alphaIIbbeta3 integrin and protein kinase C, translocation to the cytoskeleton and
association with Shc through Grb2. Biochem J 347:561–569, 2000.
286. Okamoto M, Yamamoto T, Matsubara S, Kukita I, Takeya M, Miyauchi Y, Kambara T.
Factor XIII-dependent generation of 5th complement component(C5)-derived
monocyte chemotactic factor coinciding with plasma clotting. Biochim Biophys Acta
1138:53– 61, 1992.
287. Ortner E, Schroeder V, Walser R, Zerbe O, Kohler HP. Sensitive and selective detection of free FXIII activation peptide: a potential marker of acute thrombotic events.
Blood 115:5089 –5096, 2010.
288. Otto HH, Schirmeister T. Cysteine proteases and their inhibitors. Chem Rev 97:133–
172, 1997.
289. Ozsoyl US, Hicsonmez G. Platelet aggregation in congenital factor XIII deficiency.
Acta Haematol 56:314 –317, 1976.
290. Pabst MJ, Pabst KM, Handsman DB, Beranova-Giorgianni S, Giorgianni F. Proteome
of monocyte priming by lipopolysaccharide, including changes in interleukin-1beta
and leukocyte elastase inhibitor. Proteome Sci 6:13, 2008.
305. Radek JT, Jeong JM, Wilson J, Lorand L. Association of the A subunits of recombinant
placental factor XIII with the native carrier B subunits from human plasma. Biochemistry 32:3527–3534, 1993.
306. Rao KM, Newcomb TF. Clot retraction in a factor XIII free system. Scand J Haematol
24:142–148, 1980.
307. Reed GL, Houng AK. The contribution of activated factor XIII to fibrinolytic resistance
in experimental pulmonary embolism. Circulation 99:299 –304, 1999.
308. Reed GL, Matsueda GR, Haber E. Platelet factor XIII increases the fibrinolytic resistance of platelet-rich clots by accelerating the crosslinking of alpha 2-antiplasmin to
fibrin. Thromb Haemost 68:315–320, 1992.
309. Rex S, Beaulieu LM, Perlman DH, Vitseva O, Blair PS, McComb ME, Costello CE,
Freedman JE. Immune versus thrombotic stimulation of platelets differentially regulates signalling pathways, intracellular protein-protein interactions, and alpha-granule
release. Thromb Haemost 102:97–110, 2009.
310. Rider DM, McDonagh J. Resistance of factor XIII to degradation or activation by
plasmin. Biochim Biophys Acta 675:171–177, 1981.
311. Rijken DC, Lijnen HR. New insights into the molecular mechanisms of the fibrinolytic
system. J Thromb Haemost 7:4 –13, 2009.
312. Ritchie H, Lawrie LC, Crombie PW, Mosesson MW, Booth NA. Cross-linking of
plasminogen activator inhibitor 2 and alpha 2-antiplasmin to fibrin(ogen). J Biol Chem
275:24915–24920, 2000.
291. Park D, Choi SS, Ha KS. Transglutaminase 2: a multi-functional protein in multiple
subcellular compartments. Amino Acids 39:619 – 631, 2010.
313. Ritchie H, Robbie LA, Kinghorn S, Exley R, Booth NA. Monocyte plasminogen activator inhibitor 2 (PAI-2) inhibits u-PA-mediated fibrin clot lysis and is cross-linked to
fibrin. Thromb Haemost 81:96 –103, 1999.
292. Patthy L. Detecting homology of distantly related proteins with consensus sequences.
J Mol Biol 198:567–577, 1987.
314. Robbins KC. A study on the conversion of fibrinogen to fibrin. Am J Physiol 142:581–
588, 1944.
293. Patthy L. Evolution of the proteases of blood coagulation and fibrinolysis by assembly
from modules. Cell 41:657– 663, 1985.
294. Pedersen LC, Yee VC, Bishop PD, Le Trong I, Teller DC, Stenkamp RE. Transglutaminase factor XIII uses proteinase-like catalytic triad to crosslink macromolecules.
Protein Sci 3:1131–1135, 1994.
295. Peng X, Zhang Y, Zhang H, Graner S, Williams JF, Levitt ML, Lokshin A. Interaction of
tissue transglutaminase with nuclear transport protein importin-alpha3. FEBS Lett
446:35–39, 1999.
296. Penzes K, Kover KE, Fazakas F, Haramura G, Muszbek L. Molecular mechanism of the
interaction between activated factor XIII and its glutamine donor peptide substrate. J
Thromb Haemost 7:627– 633, 2009.
970
315. Robinson BR, Houng AK, Reed GL. Catalytic life of activated factor XIII in thrombi.
Implications for fibrinolytic resistance and thrombus aging. Circulation 102:1151–
1157, 2000.
316. Rodeghiero F, Castaman GC, Di Bona E, Ruggeri M, Dini E. Successful pregnancy in a
woman with congenital factor XIII deficiency treated with substitutive therapy. Report of a second case. Blut 55:45– 48, 1987.
317. Romanic AM, Arleth AJ, Willette RN, Ohlstein EH. Factor XIIIa cross-links lipoprotein(a) with fibrinogen and is present in human atherosclerotic lesions. Circ Res 83:
264 –269, 1998.
318. Rosenthal AK, Masuda I, Gohr CM, Derfus BA, Le M. The transglutaminase, Factor
XIIIA, is present in articular chondrocytes. Osteoarthritis Cartilage 9:578 –581, 2001.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
281. Nurminskaya MV, Linsenmayer TF. Immunohistological analysis of transglutaminase
factor XIIIA expression in mouse embryonic growth plate. J Orthop Res 20:575–578,
2002.
MULTIPLE FUNCTIONS OF FACTOR XIII
319. Rosenthal AK, Mosesson MW, Gohr CM, Masuda I, Heinkel D, Seibenlist KR. Regulation of transglutaminase activity in articular chondrocytes through thrombin receptor-mediated factor XIII synthesis. Thromb Haemost 91:558 –568, 2004.
320. Roth WJ, Fleit HB, Chung SI, Janoff A. Characterization of two distinct transglutaminases of murine bone marrow-derived macrophages: effects of exposure of viable
cells to cigarette smoke on enzyme activity. J Leukoc Biol 42:9 –20, 1987.
321. Ryan AW, Hughes DA, Tang K, Kelleher DP, Ryan T, McManus R, Stoneking M.
Natural selection and the molecular basis of electrophoretic variation at the coagulation F13B locus. Eur J Hum Genet 17:219 –227, 2009.
322. Sabo TM, Brasher PB, Maurer MC. Perturbations in factor XIII resulting from activation and inhibition examined by solution based methods and detected by MALDI-TOF
MS. Biochemistry 46:10089 –10101, 2007.
323. Saito M, Asakura H, Yoshida T, Ito K, Okafuji K, Matsuda T. A familial factor XIII
subunit B deficiency. Br J Haematol 74:290 –294, 1990.
324. Sakata Y, Aoki N. Cross-linking of alpha 2-plasmin inhibitor to fibrin by fibrin-stabilizing factor. J Clin Invest 65:290 –297, 1980.
326. Sakata Y, Mimuro J, Aoki N. Differential binding of plasminogen to crosslinked and
noncrosslinked fibrins: its significance in hemostatic defect in factor XIII deficiency.
Blood 63:1393–1401, 1984.
327. Sanchez C, Deberg MA, Bellahcene A, Castronovo V, Msika P, Delcour JP, Crielaard
JM, Henrotin YE. Phenotypic characterization of osteoblasts from the sclerotic zones
of osteoarthritic subchondral bone. Arthritis Rheum 58:442– 455, 2008.
328. Sane DC, Moser TL, Pippen AM, Parker CJ, Achyuthan KE, Greenberg CS. Vitronectin is a substrate for transglutaminases. Biochem Biophys Res Commun 157:115–120,
1988.
329. Sarvary A, Szucs S, Balogh I, Becsky A, Bardos H, Kavai M, Seligsohn U, Egbring R,
Lopaciuk S, Muszbek L, Adany R. Possible role of factor XIII subunit A in Fcgamma and
complement receptor-mediated phagocytosis. Cell Immunol 228:81–90, 2004.
330. Schroder V, Kohler HP. Effect of factor XIII Val34Leu on alpha2-antiplasmin incorporation into fibrin. Thromb Haemost 84:1128 –1130, 2000.
331. Schroeder V, Durrer D, Meili E, Schubiger G, Kohler HP. Congenital factor XIII
deficiency in Switzerland: from the worldwide first case in 1960 to its molecular
characterisation in 2005. Swiss Med Wkly 137:272–278, 2007.
341. Shemirani AH, Haramura G, Bagoly Z, Muszbek L. The combined effect of fibrin
formation and factor XIII A subunit Val34Leu polymorphism on the activation of factor
XIII in whole plasma. Biochim Biophys Acta 1764:1420 –1423, 2006.
342. Siebenlist KR, Meh DA, Mosesson MW. Protransglutaminase (factor XIII) mediated
crosslinking of fibrinogen and fibrin. Thromb Haemost 86:1221–1228, 2001.
343. Siebenlist KR, Mosesson MW. Evidence of intramolecular cross-linked A alpha.gamma
chain heterodimers in plasma fibrinogen. Biochemistry 35:5817–5821, 1996.
344. Siebenlist KR, Mosesson MW. Progressive cross-linking of fibrin gamma chains increases resistance to fibrinolysis. J Biol Chem 269:28414 –28419, 1994.
345. Siebenlist KR, Mosesson MW, Hernandez I, Bush LA, Di Cera E, Shainoff JR, Di Orio
JP, Stojanovic L. Studies on the basis for the properties of fibrin produced from
fibrinogen-containing gamma= chains. Blood 106:2730 –2736, 2005.
346. Sixma JJ, van den Berg A, Schiphorst M, Geuze HJ, McDonagh J. Immunocytochemical
localization of albumin and factor XIII in thin cryo sections of human blood platelets.
Thromb Haemost 51:388 –391, 1984.
347. Skorstengaard K, Halkier T, Hojrup P, Mosher D. Sequence location of a putative
transglutaminase cross-linking site in human vitronectin. FEBS Lett 262:269 –274,
1990.
348. Sobel JH, Gawinowicz MA. Identification of the alpha chain lysine donor sites involved
in factor XIIIa fibrin cross-linking. J Biol Chem 271:19288 –19297, 1996.
349. Sorensen ES, Rasmussen LK, Moller L, Jensen PH, Hojrup P, Petersen TE. Localization
of transglutaminase-reactive glutamine residues in bovine osteopontin. Biochem J 304:
13–16, 1994.
350. Souri M, Kaetsu H, Ichinose A. Sushi domains in the B subunit of factor XIII responsible
for oligomer assembly. Biochemistry 47:8656 – 8664, 2008.
351. Souri M, Koseki-Kuno S, Takeda N, Degen JL, Ichinose A. Administration of factor XIII
B subunit increased plasma factor XIII A subunit levels in factor XIII B subunit knockout mice. Int J Hematol 87:60 – 68, 2008.
352. Souri M, Koseki-Kuno S, Takeda N, Yamakawa M, Takeishi Y, Degen JL, Ichinose A.
Male-specific cardiac pathologies in mice lacking either the A or B subunit of factor
XIII. Thromb Haemost 99:401– 408, 2008.
332. Schroeder V, Vuissoz JM, Caflisch A, Kohler HP. Factor XIII activation peptide is
released into plasma upon cleavage by thrombin and shows a different structure
compared to its bound form. Thromb Haemost 97:890 – 898, 2007.
353. Standeven KF, Carter AM, Grant PJ, Weisel JW, Chernysh I, Masova L, Lord ST, Ariens
RA. Functional analysis of fibrin ␥-chain cross-linking by activated factor XIII: determination of a cross-linking pattern that maximizes clot stiffness. Blood 110:902–907,
2007.
333. Schroth M, Meissner U, Cesnjevar R, Weyand M, Singer H, Rascher W, Klinge J.
Plasmatic factor XIII reduces severe pleural effusion in children after open-heart
surgery. Pediatr Cardiol 27:56 – 60, 2006.
354. Stirling Y, Woolf L, North WR, Seghatchian MJ, Meade TW. Haemostasis in normal
pregnancy. Thromb Haemost 52:176 –182, 1984.
334. Schwartz ML, Pizzo SV, Hill RL, McKee PA. Human Factor XIII from plasma and
platelets. Molecular weights, subunit structures, proteolytic activation, and crosslinking of fibrinogen and fibrin. J Biol Chem 248:1395–1407, 1973.
355. Streit M, Velasco P, Riccardi L, Spencer L, Brown LF, Janes L, Lange-Asschenfeldt B,
Yano K, Hawighorst T, Iruela-Arispe L, Detmar M. Thrombospondin-1 suppresses
wound healing and granulation tissue formation in the skin of transgenic mice. EMBO
J 19:3272–3282, 2000.
335. Seelig GF, Folk JE. Noncatalytic subunits of human blood plasma coagulation factor
XIII. Preparation and partial characterization of modified forms. J Biol Chem 255:
8881– 8886, 1980.
336. Seitz R, Duckert F, Lopaciuk S, Muszbek L, Rodeghiero F, Seligsohn U. ETRO Working Party on Factor XIII questionnaire on congenital factor XIII deficiency in Europe:
status and perspectives. Study Group Semin Thromb Hemost 22:415– 418, 1996.
337. Seitz R, Leugner F, Katschinski M, Immel A, Kraus M, Egbring R, Goke B. Ulcerative
colitis and Crohn’s disease: factor XIII, inflammation and haemostasis. Digestion 55:
361–367, 1994.
338. Seiving B, Ohlsson K, Linder C, Stenberg P. Transglutaminase differentiation during
maturation of human blood monocytes to macrophages. Eur J Haematol 46:263–271,
1991.
339. Semba U, Chen J, Ota Y, Jia N, Arima H, Nishiura H, Yamamoto T. A plasma protein
indistinguishable from ribosomal protein S19: conversion to a monocyte chemotactic
factor by a factor XIIIa-catalyzed reaction on activated platelet membrane phosphatidylserine in association with blood coagulation. Am J Pathol 176:1542–1551, 2010.
356. Sugimura Y, Hosono M, Wada F, Yoshimura T, Maki M, Hitomi K. Screening for the
preferred substrate sequence of transglutaminase using a phage-displayed peptide
library: identification of peptide substrates for TGASE 2 and Factor XIIIA. J Biol Chem
281:17699 –17706, 2006.
357. Sumi Y, Ichikawa Y, Nakamura Y, Miura O, Aoki N. Expression and characterization
of pro alpha 2-plasmin inhibitor. J Biochem 106:703–707, 1989.
358. Swirski FK, Nahrendorf M, Etzrodt M, Wildgruber M, Cortez-Retamozo V, Panizzi P,
Figueiredo JL, Kohler RH, Chudnovskiy A, Waterman P, Aikawa E, Mempel TR, Libby
P, Weissleder R, Pittet MJ. Identification of splenic reservoir monocytes and their
deployment to inflammatory sites. Science 325:612– 616, 2009.
359. Takahashi N, Takahashi Y, Putnam FW. Primary structure of blood coagulation factor
XIIIa (fibrinoligase, transglutaminase) from human placenta. Proc Natl Acad Sci USA
83:8019 – 8023, 1986.
360. Tamaki T, Aoki N. Cross-linking of alpha 2-plasmin inhibitor to fibrin catalyzed by
activated fibrin-stabilizing factor. J Biol Chem 257:14767–14772, 1982.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
971
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
325. Sakata Y, Aoki N. Significance of cross-linking of alpha 2-plasmin inhibitor to fibrin in
inhibition of fibrinolysis and in hemostasis. J Clin Invest 69:536 –542, 1982.
340. Serrano K, Devine DV. Intracellular factor XIII crosslinks platelet cytoskeletal elements upon platelet activation. Thromb Haemost 88:315–320, 2002.
MUSZBEK ET AL.
376. Weisel JW, Litvinov RI. The biochemical and physical process of fibrinolysis and effects
of clot structure and stability on the lysis rate. Cardiovasc Hematol Agents Med Chem
6:161–180, 2008.
362. Torocsik D, Bardos H, Nagy L, Adany R. Identification of factor XIII-A as a marker of
alternative macrophage activation. Cell Mol Life Sci 62:2132–2139, 2005.
377. Weiss MS, Metzner HJ, Hilgenfeld R. Two non-proline cis peptide bonds may be
important for factor XIII function. FEBS Lett 423:291–296, 1998.
363. Torocsik D, Szeles L, Paragh G ,Jr, Rakosy Z, Bardos H, Nagy L, Balazs M, Inbal A,
Adany R. Factor XIII-A is involved in the regulation of gene expression in alternatively
activated human macrophages. Thromb Haemost 104:709 –717, 2010.
378. Wells PS, Anderson JL, Scarvelis DK, Doucette SP, Gagnon F. Factor XIII Val34Leu
variant is protective against venous thromboembolism: a huge review and metaanalysis. Am J Epidemiol 164:101–109, 2006.
364. Turner BT ,Jr, Maurer MC. Evaluating the roles of thrombin and calcium in the
activation of coagulation factor XIII using H/D exchange and MALDI-TOF MS. Biochemistry 41:7947–7954, 2002.
379. Wolpl A, Lattke H, Board PG, Arnold R, Schmeiser T, Kubanek B, Robin-Winn M,
Pichelmayr R, Goldmann SF. Coagulation factor XIII A and B subunits in bone marrow
and liver transplantation. Transplantation 43:151–153, 1987.
365. Turner BT ,Jr, Sabo TM, Wilding D, Maurer MC. Mapping of factor XIII solvent
accessibility as a function of activation state using chemical modification methods.
Biochemistry 43:9755–9765, 2004.
380. Wozniak G, Dapper F, Alemany J. Factor XIII in ulcerative leg disease: background and
preliminary clinical results. Semin Thromb Hemost 22:445– 450, 1996.
366. Ueyama M, Urayama T. The role of factor XIII in fibroblast proliferation. Jpn J Exp Med
48:135–142, 1978.
381. Wozniak G, Noll T, Akinturk H, Thul J, Muller M. Factor XIII prevents development of
myocardial edema in children undergoing surgery for congenital heart disease. Ann NY
Acad Sci 936:617– 620, 2001.
367. Valnickova Z, Enghild JJ. Human procarboxypeptidase U, or thrombin-activable fibrinolysis inhibitor, is a substrate for transglutaminases. Evidence for transglutaminasecatalyzed cross-linking to fibrin. J Biol Chem 273:27220 –27224, 1998.
382. Yee VC, Pedersen LC, Bishop PD, Stenkamp RE, Teller DC. Structural evidence that
the activation peptide is not released upon thrombin cleavage of factor XIII. Thromb
Res 78:389 –397, 1995.
368. Van Wersch JW, Vooijs ME, Ubachs JM. Coagulation factor XIII in pregnant smokers
and non-smokers. Int J Clin Lab Res 27:68 –71, 1997.
383. Yee VC, Pedersen LC, Le Trong I, Bishop PD, Stenkamp RE, Teller DC. Threedimensional structure of a transglutaminase: human blood coagulation factor XIII. Proc
Natl Acad Sci USA 91:7296 –7300, 1994.
369. Vanscheidt W, Hasler K, Wokalek H, Niedner R, Schopf E. Factor XIII-deficiency in
the blood of venous leg ulcer patients. Acta Derm Venereol 71:55–57, 1991.
370. Voko Z, Bereczky Z, Katona E, Adany R, Muszbek L. Factor XIII Val34Leu variant
protects against coronary artery disease. A meta-analysis. Thromb Haemost 97:458 –
463, 2007.
371. Walter M, Nyman D, Krajnc V, Duckert F. The activation of plasma factor XIII with the
snake venom enzymes ancrod and batroxobin marajoensis. Thromb Haemost 38:438 –
446, 1977.
372. Wang DL, Annamalai AE, Ghosh S, Gewirtz AM, Colman RW. Human platelet factor
V is crosslinked to actin by FXIIIa during platelet activation by thrombin. Thromb Res
57:39 –57, 1990.
373. Wang G, Beier F. Rac1/Cdc42 and RhoA GTPases antagonistically regulate chondrocyte proliferation, hypertrophy, and apoptosis. J Bone Miner Res 20:1022–1031, 2005.
374. Wartiovaara U, Mikkola H, Szoke G, Haramura G, Karpati L, Balogh I, Lassila R,
Muszbek L, Palotie A. Effect of Val34Leu polymorphism on the activation of the
coagulation factor XIII-A. Thromb Haemost 84:595– 600, 2000.
375. Webb GC, Coggan M, Ichinose A, Board PG. Localization of the coagulation factor
XIII B subunit gene (F13B) to chromosome bands 1q31–32.1 and restriction fragment
length polymorphism at the locus. Hum Genet 81:157–160, 1989.
972
384. Yorifuji H, Anderson K, Lynch GW, Van de Water L, McDonagh J. B protein of factor
XIII: differentiation between free B and complexed B. Blood 72:1645–1650, 1988.
385. Zamboni P, De Mattei M, Ongaro A, Fogato L, Carandina S, De Palma M, Tognazzo
S, Scapoli GL, Serino ML, Caruso A, Liboni A, Gemmati D. Factor XIII contrasts the
effects of metalloproteinases in human dermal fibroblast cultured cells. Vasc Endovascular Surg 38:431– 438, 2004.
386. Zamboni P, Gemmati D. Clinical implications of gene polymorphisms in venous leg
ulcer: a model in tissue injury and reparative process. Thromb Haemost 98:131–137,
2007.
387. Zhang R, Murakami S, Coustry F, Wang Y, de Crombrugghe B. Constitutive activation
of MKK6 in chondrocytes of transgenic mice inhibits proliferation and delays endochondral bone formation. Proc Natl Acad Sci USA 103:365–370, 2006.
388. Zhou J, Tang PC, Qin L, Gayed PM, Li W, Skokos EA, Kyriakides TR, Pober JS, Tellides
G. CXCR3-dependent accumulation and activation of perivascular macrophages is
necessary for homeostatic arterial remodeling to hemodynamic stresses. J Exp Med
207:1951–1966, 2010.
389. Zhu Y, Tassi L, Lane W, Mendelsohn ME. Specific binding of the transglutaminase,
platelet factor XIII, to HSP27. J Biol Chem 269:22379 –22384, 1994.
Physiol Rev • VOL 91 • JULY 2011 • www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on June 14, 2017
361. Tognazzo S, Gemmati D, Palazzo A, Catozzi L, Carandina S, Legnaro A, Tacconi G,
Scapoli GL, Zamboni P. Prognostic role of factor XIII gene variants in nonhealing
venous leg ulcers. J Vasc Surg 44:815– 819, 2006.