ArticlePDF AvailableLiterature Review

Chemokine and cytokine processing by matrix metalloproteinases and its effect on leukocyte migration and inflammation

Authors:

Abstract

The action of matrix metalloproteinases (MMPs) was originally believed to be restricted to degradation of the extracellular matrix; however, in recent years, it has become evident that these proteases can modify many nonmatrix substrates, such as cytokines and chemokines. The use of MMP-deficient animals has revealed that these proteases can indeed influence the progression of various inflammatory processes. This review aims to provide the reader with a concise overview of these novel MMP functions in relation to leukocyte migration.
Chemokine and cytokine processing by matrix
metalloproteinases and its effect on leukocyte migration
and inflammation
Philippe Van Lint
1
and Claude Libert
2
Departments of Molecular Biomedical Research, VIB, and Molecular Biology, Ghent University, Ghent, Belgium
Abstract: The action of matrix metalloprotein-
ases (MMPs) was originally believed to be restricted
to degradation of the extracellular matrix; how-
ever, in recent years, it has become evident that
these proteases can modify many nonmatrix sub-
strates, such as cytokines and chemokines. The use
of MMP-deficient animals has revealed that these
proteases can indeed influence the progression of
various inflammatory processes. This review aims
to provide the reader with a concise overview of
these novel MMP functions in relation to leukocyte
migration. J. Leukoc. Biol. 82: 000 000; 2007.
Key Words: extracellular matrix ELR
INTRODUCTION
Matrix metalloproteinases (MMPs) form a family of closely
related, zinc-dependent endoproteinases. The first report about
MMPs dates back to 1962, when Gross and Lapiere [1], while
attempting to establish how a tadpole loses its tail during
metamorphosis, discovered the first member of this family
(MMP-1). Since then, the family has expanded gradually; the
latest member discovered is MMP-28 [2], which gives a total of
23 human MMPs. As their name suggests, MMPs were char-
acterized initially as matrix-degrading proteases. Indeed, col-
lectively, these enzymes can degrade all components of the
extracellular matrix (ECM), thereby influencing many impor-
tant processes, such as cell proliferation, differentiation, mi-
gration, and death, as well as cell– cell interactions [3]. It is
therefore not surprising that MMPs play a crucial role in many
physiological processes, e.g., bone morphogenesis, the men-
strual cycle, and development and also, in many pathological
conditions, such as cancer invasion, arthritis, and atheroscle-
rosis. The story has been made even more complex by the
increasing number of studies revealing many nonmatrix sub-
strates for MMPs, such as chemokines, growth factors, and
receptors, indicating that MMPs influence an even wider array
of physiological and pathological processes [4].
Inflammatory conditions are almost always characterized by
deregulated, often increased MMP activities [5]. The central
question is whether these MMPs can influence the outcome of
inflammation and if so, how they do so. Only by understanding
the mechanism by which MMPs exert their function can they
develop into effective drug targets. Indeed, ignorance of the
relevant in vivo substrates is one of the main reasons for the
disappointing and often unexpected results of trials using MMP
inhibitors. Identifying the functions of specific MMPs in spe-
cific pathologies should therefore become a major goal. Ini-
tially, knowledge of MMP substrates was based on in vitro
experiments, in which purified, activated MMPs were incu-
bated with specific substrates. Although these data can be
informative, the question remains whether processing of a
purified substrate under optimal conditions implies that the
same protein is a physiologically relevant MMP substrate in
vivo. In this respect, the use of transgene and knockout tech-
nologies has meant a giant leap forward, as it enables the
testing of MMP functions in vivo.
Apart from occasional, subtle developmental differences,
such as a transient delay in myelination observed in MMP-9
and MMP-12 knockout mice [6], all MMP-deficient animals
have an overall normal development and are viable. A notable
exception is the MT1-MMP (MMP-14) knockout, which dis-
plays severe skeletal deformations and dies shortly after birth
[7, 8]. However, many MMP knockouts challenged by injury,
inflammatory stimulus, infection, or cancer reveal interesting
phenotypes. Indeed, numerous studies have reported that
MMPs play roles in modulating inflammatory reactions by
acting at different levels: leukocyte recruitment, alteration of
the functions of cytokines and proteases, and clearance of the
pathogen [5].
MMPs AND LEUKOCYTE RECRUITMENT
Leukocyte recruitment to the site of injury or infection is a
complex process, which is regulated tightly by the interplay
between endothelial cells and leukocytes, a process in which
chemokines play a central role. These chemotactic cytokines
together help attract leukocytes to the site of infection or
injury, thereby influencing the outcome of an inflammatory
response. They are divided into four structural groups: CC
chemokines (or -chemokines) have two adjacent cysteines
1
Current address: General Hospital Sint-Augustinus, Clinical Laboratory,
Wilrijk, Belgium.
2
Correspondence: VIB, Ghent University, Technologiepark 927, Ghent (Zwi-
jnaarde), East-Flanders 9052, Belgium. E-mail: Claude.Libert@dmbr.Ugent.be
Received June 1, 2007; revised July 19, 2007; accepted July 20, 2007.
doi: 10.1189/jlb.0607338
0741-5400/07/0082-0001 © Society for Leukocyte Biology Journal of Leukocyte Biology Volume 82, December 2007 1
Uncorrected Version. Published on August 20, 2007 as DOI:10.1189/jlb.0607338
Copyright 2007 by The Society for Leukocyte Biology.
near the N terminus; CXC chemokines (or -chemokines), in
which the cysteines are separated by a variable amino acid;
CX
3
C chemokines (or -chemokines) have three variable
amino acids between the two cysteines; and C chemokines (or
-chemokines) have only one cysteine at the N terminus [9].
Another important structural property of chemokines is the
presence or absence of a specific tripeptide sequence Glu-Leu-
Arg (ELR), which is important in the interaction with CXCR1
and CXCR2 [10].
Chemokine processing by MMPs
As shown in Table 1, MMP-mediated proteolysis of chemo-
kines is one way by which MMPs can influence leukocyte
trafficking. MMP proteolysis can affect the biological functions
of chemokines in different ways. First, the proteolysis might
inactivate the chemokine. Second, processing might generate
antagonistic derivatives, which can still bind to the chemokine
receptor but cannot promote chemotaxis. Third, the truncated
chemokine is a more powerful chemotactic agent. Whatever the
outcome, chemokine processing undoubtedly affects the pro-
gression of an inflammatory reaction and influences the type of
cells, which are recruited and activated.
It was shown that MMP-2 can cleave CXCL12 (also known
as stromal cell-derived factor-1) to produce a truncated pro-
tein, which appeared to be highly neurotoxic, although it did
not affect chemotaxis directly [11]. CXCL12 inactivation is not
restricted to MMP-2, as MMP-1, MMP-3, MMP-9, MMP-13,
and MMP-14 have been shown to inactivate this chemokine
[12]. MMP-2 was also found to shed the plasma membrane-
associated chemokine CX
3
CL1 (fractalkine) to generate a
soluble chemokine; however, an additional MMP-2-medi-
ated cleavage at the N terminus of the protein inactivates
the chemokine, converting it into a potent, functional an-
tagonist [13].
MMP-9 also inactivates CXCL chemokines. For instance,
CXCL4 (platelet factor 4), a ELR-negative CXC chemokine,
and CXCL1 (growth-related oncogene-), a ELR motif contain-
ing CXC chemokine [14], are proteolytically inactivated by
MMP-9. This MMP also degraded CTAPIII, which is the NH2-
terminally extended, inactive precursor of CXCL7 (neutrophil-
activating peptide-2 [14]). CXCL5 [epithelial-derived neutro-
phil-activating factor-78 (ENA-78)] is also cleaved by gelati-
nase B at different sites, resulting in transient potentiation of
this chemokine, but eventually leading to its inactivation [15].
Finally, CXCL9 (monokine induced by IFN-) and CXCL10
(IFN-inducible protein 10) are processed C-terminally by
MMP-8 and MMP-9 [16]. At least in the case of CXCL9, this
decreases chemotactic ability drastically and might intervene
with association with the ECM [17].
As mentioned earlier, several inactivated chemokines are
still capable of binding to their receptors and therefore, act as
functional antagonists, preventing the activity of other non-
cleaved chemokines. MMP-2, for instance, was shown to pro-
cess CCL7 (also known as MCP-3) into an antagonistic form
[18]. Other MMPs are also capable of cleaving MCPs to gen-
erate chemotactic antagonists. In addition to MMP-2, CCL7 is
cleaved efficiently by MMP-1, MMP-3, MMP-13, and MMP-14.
The closely related chemokines CCL2 (MCP-1) and CCL13
(MCP-4) could also be cleaved by MMP-1 and MMP-3, the
latter of which also cleaved CCL8 (MCP-2) [19]. Finally,
MMP-8 also processed CCL2, although inefficiently. It is in-
teresting that all the truncated forms generated were shown to
function as inflammatory antagonists when administered in
vivo. This means that MMPs can have anti-inflammatory effects
by dampening the action of chemokines.
In contrast to these examples, MMPs have also been shown
to increase the biological activity of chemokines. Indeed,
MMP-9 was shown to process CXCL8 (IL-8), which led to a
significant increase in its chemotactic activity [14]. MMP-8,
MMP-13, and MMP-14 were later also shown to generate a
truncated IL-8 species with increased activity [20, 21]. More-
over, another human CXCR2 ligand, CXCL5 (ENA-78), was
found recently to be activated by MMP-8 cleavage [21].
The rodent chemokine LIX, which is considered to be the
sole murine counterpart of two closely related human chemo-
kines, namely CXCL5 (ENA-78) and CXCL6 [granulocyte che-
motactic protein 2 (GCP-2)], was found to be processed N-
terminally by MMP-9, with a consequent twofold increase in its
biological activity [15]. MMP-8 was later shown to cleave LIX
N- and C-terminally, generating truncated forms with in-
creased chemotactic activities [22]. Recently, MMP-1, MMP-2,
and MMP-13 were also shown to enhance the biological func-
tion of LIX by processing it N-terminally [21]. These observa-
tions of MMP-mediated LIX activation can have a serious
impact on the course of several neutrophil-driven pathologies,
as mouse LIX is believed to play the same role as IL-8 in
TABLE 1. Chemokine Processing by MMPs and Its Effect on Chemotactic Capacity
LIX CX3CL1 CXCL1 CXCL4 CXCL5 CXCL6
CTAPIII
(CXCL7) CXCL8 CXCL9 CXCL10 CXCL12 CCL2 CCL7 CCL8 CCL13
MMP-1 ⫹⫹⫹ ??—⫺⫺⫺ ⫺⫺⫺ ⫺⫺⫺
MMP-2 ⫹⫹⫹ ⫺⫺⫺ ⫺⫺⫺
MMP-3 ⫺⫺⫺ ⫺⫺⫺ ⫺⫺⫺ ⫺⫺⫺
MMP-8 ⫹⫹⫹ ⫹⫹⫹ 0⫹⫹⫹ —? ⫺⫺⫺
MMP-9 ⫹⫹⫹ ——— 0 ⫹⫹⫹ —?
MMP-13 ⫹⫹⫹ ⫹⫹⫹ ⫺⫺⫺
MMP-14 ⫹⫹⫹ ⫺⫺⫺
Proteolytic modification of chemokines by MMPs can have different consequences for the chemokines biological activity. For some chemokines, the effect of
proteolysis on their chemotactic capacity was not reported (?), and other chemokines have been shown to become inactivated (–) or transformed into an antagonist
(⫺⫺⫺). Proteolysis can also lead to an increase in chemotactic potency of the chemotactic protein (⫹⫹⫹) or have no effect on chemokine activity (0). Gray boxes
indicate that proteolysis has not been reported. LIX, LPS-induced CXC chemokine; CTAPIII, connective tissue-activating peptide III.
2 Journal of Leukocyte Biology Volume 82, December 2007 http://www.jleukbio.org
humans. Indeed, compared with humans, rodents have few
chemokines. For instance, the mouse has only four ELR-
containing CXC chemokines, which specifically target poly-
morphonuclear cells: LIX, keratinocyte-derived chemokine
(KC), dendritic cell (DC) inflammatory protein-1, and MIP-2, of
which only LIX was found to be processed by MMPs [21].
Finally, it is noteworthy that cleavage does not always alter
the activities of chemokines. An example of this is the N-
terminal cleavage of CXCL6 (GCP-2) by MMP-8 and MMP-9,
which does not affect its biological activity [15].
MMPs modulating chemokine gradients
In vivo, however, instead of being presented to proteases as
soluble proteins, chemokines are immobilized mostly on the
ECM or cell surface by binding to glycosaminoglycans (GAGs)
through positively charged domains. This binding to GAGs
seems to be important for chemokines to exert their role
effectively. IL-8 binding to heparan sulfate leads to structural
stabilization of the dimeric form of the chemokine, resulting in
an extended biological activity and enhanced neutrophil re-
sponses [23]. Some chemokines with mutations in the GAG-
binding sites lose the ability to recruit cells in vivo, but they
retain receptor-signaling activity in vitro [24]. This might mean
that MMPs mediate the function of chemokines indirectly by
releasing chemokines bound to the cell surface or ECM.
The best example of this comes from a study describing the
role of MMP-7 in a bleomycin-induced model of lung inflam-
mation [25]. In response to the mucosal damage, epithelial
cells release KC, which upon secretion, binds to syndecan-1, a
cell-bound heparan sulfate proteoglycan (HSPG). It was al-
ready known that a tissue inhibitor of metalloproteinase 3
(TIMP-3)-sensitive protease could mediate the cleavage of the
intact syndecan-1 and syndecan-4 ectodomains (shedding),
thereby converting the cell-surface molecules into soluble ef-
fectors [26]. Indeed, together with CXCL1, MMP-7 is produced
by these epithelial cells and cleaves the ectodomain of synde-
can-1, thereby releasing the syndecan-1/KC complex. This
creates a chemokine gradient, which triggers neutrophil infil-
tration into the alveolar space. As seen after bleomycin injury,
neutrophils of the MMP-7-deficient animals do extravasate, but
in contrast to their wild-type counterparts, they do not enter the
lumen of the lung. Instead, they are confined to an expanded
perivascular space. As a consequence, survival of MMP-7
knockouts was much better than that of the wild-types. It is
likely that MMP-7 itself first binds to syndecan-1, as MMP-7
has been known to interact with GAGs, which markedly en-
hanced the activity of MMP-7 [27]. Therefore, interaction with
the GAG side-chains of syndecan-1 would make it easier for
MMP-7 to cleave the core protein of this HSPG.
Other experiments have confirmed that MMP-7 is indeed
needed to generate a chemokine gradient, rather than being
indispensable for neutrophil migration as such. Instillation of a
bacterial chemotactic protein in the lungs of MMP-7-deficient
mice did trigger neutrophils to infiltrate the luminal tissue, and
the outcome was worse in these animals than in their wild-type
counterparts [25].
Syndecan shedding is not restricted to MMP-7 and can be
performed by various proteases, including other MMPs.
MMP-9 can shed syndecan-1 and syndecan-4 [28], and MT1-
MMP and MT3-MMP can release the ectodomain of synde-
can-1 [29]. As chemokines, as well as many cytokines and
antimicrobial substances, have the ability to bind to GAGs, this
protease-controlled release of syndecans can affect the pro-
gression of different pathologies.
Besides the study describing the involvement of MMP-7 in
generating a CXCL1 chemokine gradient, other studies report
involvement of MMPs in establishing chemokine gradients.
Our own research has shown that in the liver, following TNF-
induced hepatitis, MMP-8 appears to be indispensable for the
release of yet another neutrophil chemoattractant, LIX [30]. As
a result, neutrophil influx is seriously impaired in mice lacking
a functional Mmp8 gene, which improves survival dramatically.
Moreover, in a model of allergen-induced asthma, leukocytes
accumulated massively in the lung parenchyma but in contrast
to wild-type animals, failed to reach the airway lumen in
MMP-2-deficient animals [31]. This was accompanied by a
large reduction in the levels of CCL11 (eotaxin), a potent
eosinophil chemoattractant, in bronchoalveolar lavage (BAL)
fluids of MMP-2 knockouts. The same group extended the
study by applying the model to MMP-9 knockouts and MMP-
2/9 double-knockouts [32]. In contrast to the results obtained
with the MMP-2 knockouts, in which the reduction in luminal
leukocytes could be attributed entirely to reduced eosinophil
numbers, MMP-9- and MMP-2/9-deficient animals had fewer
eosinophils and neutrophils in their BAL. Furthermore, al-
though absence of MMP-2 affected only CCL11 levels, lack of
MMP-9 led to significantly lower levels of CCL11, CCL7
(MCP-3), and CCL17 (thymus and activation-regulated chemo-
kine). Unfortunately, the mechanism by which both gelatinases
modulate levels of these chemokines in BAL is still unclear.
The fact that MMP-9 seems to promote neutrophil migration in
this model is interesting in view of another study by Ste-
fanidakis et al. [33]. They found that MMP-9 forms a complex
with the
M
2
integrin in the intracellular granules of neutro-
phils and that this complex becomes localized to the cell
surface upon cellular activation. Furthermore, blocking the
interaction between MMP-9 and
M
2
inhibits leukocyte mi-
gration in vitro as in vivo, suggesting that integrin association
might help MMPs in promoting cell migration. Another study,
using a similar asthma model, also showed a marked reduction
in CCL17 BAL levels in MMP-9 knockout mice [34]. However,
this study did not report a difference in the number of recruited
eosinophils and neutrophils [35] but did show fewer DCs
migrating into the airway lumen in the absence of MMP-9 [34].
Furthermore, chondrocyte-derived MMP-3 generated an un-
identified macrophage chemotactic factor, which was required
for disc degradation in a model of herniated disc resorption
[36]. MMP-3-deficient mice also showed reduced neutrophil
recruitment to the airway lumen in a model of IgG-induced
acute lung injury, as well as decreased lung injury [37]. The
mechanism by which MMP-3 influences neutrophil migration
is unknown. However, as neutrophils are not known to produce
MMP-3, it is likely that this protease facilitates neutrophil
migration indirectly, for instance, by generating a chemokine
gradient.
MMP-12, which is predominantly a macrophage protease, is
required for the influx of these cells in a model of smoke-
induced emphysema [38]. As additional instillation of CCL2
Van Lint and Libert Matrix metalloproteinases in inflammation 3
resulted in macrophage migration comparable with that in-
duced in wild-type animals by exposure to cigarette smoke
alone, MMP-12 seems important in establishing a chemotactic
gradient, rather than in macrophage migration as such. Later,
the chemotactic proteins responsible for this phenotype were
identified as being elastin fragments (EFs) [39]. Although it has
been long known that EFs are chemotactic for monocytes in
vitro [40, 41], the study by Houghton et al. [39] is the first
report that shows that MMP-generated elastin fragments can
drive the progression of a disease. Furthermore, neutrophil
migration too seems to be driven partially by the release of
cryptic ECM fragments. Recently, it was shown that pulmonary
ECM proteolysis, following exposure to a bacterial component,
gives rise to a tripeptide chemoattractant, which promotes
neutrophil recruitment [42].
These studies are important, as they show that ECM break-
down can promote chemotaxis, not only by releasing ECM-
bound chemotactic factors but also by exposing cryptic ECM
sites possessing chemotactic properties. Indeed, after so many
chemotactic chemokine family members had been identified,
the chemotactic properties of several ECM-derived fragments
were believed to be, at best, of minimal importance in vivo.
Therefore, these data might ask for a re-evaluation of the in
vivo relevance of cryptic sites of several ECM components,
such as fibronectin, collagen, and laminin [43– 46].
Finally, nonmatrix proteins too, following proteolysis, can
give rise to unexpected chemotaxis-promoting fragments. An
example of this is the MMP-12-mediated cleavage of the 1-
proteinase inhibitor, a serine proteinase inhibitor processed by
several MMPs, which generates a fragment that promotes che-
motaxis of neutrophils [47].
MMPs and mobilization of hematopoietic
progenitor cells (HPCs)
The influence of MMPs on leukocyte migration is not limited to
the fate of mature circulating leukocytes but also affects the
trafficking of HPCs from the bone marrow into circulation.
Administrating anti-MMP-9 antibodies to rhesus monkeys to-
tally inhibited the IL-8-induced mobilization of HPCs [48].
Although a similar role for MMP-9 in IL-8-mediated HPC
release could not be shown in mice [49], MMP-9 was shown to
be involved in hematopoietic recovery after depletion of hema-
topoietic cells [50]. This could be explained by the role MMP-9
plays in shedding of the membrane-bound kit ligand, also
known as the stem cell factor. Another study indicated a
possible role for MMP-2 in releasing proteoglycan-bound
CXCL12 from the surface of bone marrow stromal cells,
thereby promoting pro-B cell migration [51].
It should be noted, however, that the influence of MMPs on
leukocyte trafficking is not merely restricted to chemokine
processing and release. Indeed, by degrading a variety of
proteins, which constitute interstitial ECMs, MMPs help leu-
kocytes cross otherwise impassable basement membranes,
such as the blood-brain barrier. For instance, MMP-2 and
MMP-9 were shown to degrade dystroglycan, a critical compo-
nent of the blood-brain barrier, thereby compromising its in-
tegrity and allowing leukocyte trafficking into the CNS during
experimental autoimmune encephalomyelitis [52].
CYTOKINE PROCESSING BY MMPs
The influence of MMPs on the progression of inflammatory
processes is not limited to leukocyte migration, as they process
not only chemokines but also a variety of cytokines. Indeed, as
with chemokines, cytokine proteolysis often leads to altered
bioavailability and activity. Shedding of the proinflammatory
cytokine TNF is one example of how MMPs might influence an
inflammatory reaction by modulating cytokines. TNF is pro-
duced as a trimeric membrane-anchored precursor and re-
leased from the cell surface by a regulated proteolytic step
[53]. In vivo studies have shown that most of the biological
functions of TNF require its shedding and release as a soluble
mediator. Indeed, mice carrying in their pro-TNF sequence a
mutation, which blocks effective release of the membrane-
anchored protein, have defective leukocyte migration [54]. A
disintegrin and metalloproteinase 17 (ADAM-17) is, as its
alternate name TNF--converting enzyme (TACE) suggests,
the main TNF sheddase. The release of active TNF is reduced
by 90% in cells from ADAM-17 knockouts, indicating that
ADAM-17 is the main TACE in vivo [55]. However, it seems
that in specific cellular settings, ADAM-17-independent re-
lease of TNF can become important. In a model of macrophage-
mediated herniated disc resorption, macrophage MMP-7 was
found to be indispensable for TNF shedding [56]. As a result,
MMP-7-deficient macrophages were unable to infiltrate the
disc. Apart from MMP-7, also MMP-1, MMP-2, MMP-3,
MMP-9, MMP-12, MMP-14, and MMP-17 were shown to re-
lease active TNF from the cell surface by a mechanism similar
to that of ADAM-17-mediated TNF release [20, 57, 58].
Another cytokine, IL-1, is produced as an inactive precur-
sor protein, which is activated by proteolytic removal of the
N-terminal part. Caspase-1, also known as IL-1-converting
enzyme (ICE), is an intracellular cysteine protease, which has
been identified as the primary IL-1activator. However, there
is evidence that IL-1can be activated in a caspase-1-inde-
pendent manner in vitro and in vivo. For instance, in response
to turpentine injection, the levels of mature IL-1were not
diminished in ICE/mice [59]. Furthermore, human kera-
tinocytes, which express IL-1, but not active ICE, were
capable of producing mature IL-1[60]. Some MMPs, namely
MMP-2, MMP-3, and MMP-9, were found to activate pro-IL-
1, but MMP-1 could not [61]. This study also showed that
further proteolytic degradation during prolonged incubation
with MMP-3 could eventually inactivate mature IL-1. Indeed,
MMP-mediated IL-1degradation had been reported already a
few years earlier by Ito et al. [62], who found that 4-amino-
phenyl mercuric acetate-activated, conditioned medium from
uterine cervical fibroblasts could degrade mature IL-1and
that this activity could be inhibited by adding TIMP-1 to the
reaction. Using purified MMP extracts, MMP-1, MMP-2,
MMP-3, and MMP-9 were shown to diminish the biological
activity of IL-1. As MMPs and IL-1 often colocalize during
inflammatory conditions, MMPs might influence positively and
negatively the inflammatory process, by activating pro-IL-1or
inactivating the mature form of this cytokine, respectively.
TGF-is an anti-inflammatory cytokine, known to restrain
the mononuclear inflammation, whose activity is tightly regu-
lated. TGF-is produced initially as a precursor protein,
4 Journal of Leukocyte Biology Volume 82, December 2007 http://www.jleukbio.org
which is cleaved in the endoplasmic reticulum by furins into an
amino-terminal fragment, called latency-associated protein,
and a shorter, carboxy-terminal fragment, which is the mature
cytokine. These fragments are assembled as a double-ho-
modimer, called the small latent complex, which is modified
further by disulfide linkage to so-called latent TGF--binding
proteins (LTBPs), thereby forming the large latent complex.
After secretion, this latent TGF-complex is cross-linked to
the ECM, forming a reservoir of latent TGF-in the extracel-
lular environment [63]. Several mechanisms, including MMP
proteolysis, have been implicated in the release of mature
TGF-from the latent complex. MMP-2 and MMP-9 [64] as
well as MMP-3 [65] and MMP-14 [66] have been identified as
TGF-activators. Moreover, MMPs have also been implicated
in the cleavage of LTBPs, thereby releasing TGF-from the
ECM [67]. MMPs can also cause release of TGF-by degrad-
ing decorin, a small, collagen-associated proteoglycan, known
to act as a depot for TGF-in the ECM [68]. If shown that these
pathways of MMP-mediated TGF-activation would be rele-
vant in vivo, this might be another mechanism by which MMP
activity restrains rather than augments inflammation. In con-
trast, MMP activity might also down-regulate TGF-signaling.
MT1-MMP was shown to shed -glycan [69], which is a mem-
brane-bound protein, functioning as a coreceptor for TGF-
and regulating the access of TGF-to its signaling receptors.
However, if released from the cell surface, at least in vitro,
-glycan functions as a TGF-inhibitor by blocking the in-
teraction between TGF-and its cell surface receptors [70].
Finally, processing and subsequent inactivation of IFN-by
MMP-9 offer another example of how MMP-mediated cytokine
proteolysis might influence the progression of an inflammatory
reaction [71].
MMPs might alter the biological activity of cytokines, not
only by direct proteolytic processing but also by shedding their
receptors. Some data suggest that MMPs contribute to the
shedding of soluble Type II IL-1 decoy receptor (sIL-1RII)
from the cell surface, as this shedding was blocked by BB-94,
a broad-spectrum MMP inhibitor [72]. The soluble receptor, by
retaining its ability to interact with IL-1, neutralizes the
biological effects of this cytokine. Thus, MMP release of sIL-
1RII offers another mechanism of how MMPs might dampen
inflammation. In contrast, other data indicate that MMPs can
degrade and therefore inactivate this soluble receptor [73].
This would mean that MMPs support rather than inhibit IL-
1-driven inflammation, but all these data have to be verified
in vivo. Other in vitro studies showed that MMP-9 cleaves
IL-2Rand thereby, down-regulates the proliferative capabil-
ity of activated T cells by generating antagonistic, sIL-2R
chains [74]. The release of the receptor of a structurally related
cytokine, namely sIL-15R, could also be blocked by using a
broad-spectrum MMP inhibitor; however, the specific MMPs
responsible remain to be identified [75].
CONCLUSION
During the last decade, it has become clear that MMPs can
influence the progression of various inflammatory conditions.
Some of the phenotypes of MMP-deficient mice could be ex-
plained by the proteolytic activities that these proteases exer-
cise in modulating the activities of various cytokines and
chemokines, as has been shown in vitro. However, although
several substrates have been identified as possible MMP tar-
gets, further identification of relevant in vivo targets is needed
for proceeding with the development of MMPs into effective
drug targets.
ACKNOWLEDGMENTS
The Institute for the Promotion of Innovation through Science
and Technology in Flanders (IWT-Vlaanderen) supported re-
search in the authors’ laboratory. The authors thank Dr. Amin
Bredan for editing this review, as well as all investigators
working in the field for their contributions.
REFERENCES
1. Gross, J., Lapiere, C. M. (1962) Collagenolytic activity in amphibian
tissues: a tissue culture assay. Proc. Natl. Acad. Sci. USA 48, 1014 –1022.
2. Lohi, J., Wilson, C. L., Roby, J. D., Parks, W. C. (2001) Epilysin, a novel
human matrix metalloproteinase (MMP-28) expressed in testis and kera-
tinocytes and in response to injury. J. Biol. Chem. 276, 10134 –10144.
3. Elkington, P. T., O’Kane, C. M., Friedland, J. S. (2005) The paradox of
matrix metalloproteinases in infectious disease. Clin. Exp. Immunol. 142,
12–20.
4. McCawley, L. J., Matrisian, L. M. (2001) Matrix metalloproteinases:
they’re not just for matrix anymore! Curr. Opin. Cell Biol. 13, 534 –540.
5. Parks, W. C., Wilson, C. L., Lopez-Boado, Y. S. (2004) Matrix metallo-
proteinases as modulators of inflammation and innate immunity. Nat. Rev.
Immunol. 4, 617– 629.
6. Larsen, P. H., DaSilva, A. G., Conant, K., Yong, V. W. (2006) Myelin
formation during development of the CNS is delayed in matrix metallo-
proteinase-9 and -12 null mice. J. Neurosci. 26, 2207–2214.
7. Holmbeck, K., Bianco, P., Caterina, J., Yamada, S., Kromer, M.,
Kuznetsov, S. A., Mankani, M., Robey, P. G., Poole, A. R., Pidoux, I.,
Ward, J. M., Birkedal-Hansen, H. (1999) MT1-MMP-deficient mice de-
velop dwarfism, osteopenia, arthritis, and connective tissue disease due to
inadequate collagen turnover. Cell 99, 81–92.
8. Zhou, Z., Apte, S. S., Soininen, R., Cao, R., Baaklini, G. Y., Rauser,
R. W., Wang, J., Cao, Y., Tryggvason, K. (2000) Impaired endochondral
ossification and angiogenesis in mice deficient in membrane-type matrix
metalloproteinase I. Proc. Natl. Acad. Sci. USA 97, 4052– 4057.
9. Zlotnik, A., Yoshie, O. (2000) Chemokines: a new classification system
and their role in immunity. Immunity 12, 121–127.
10. Loetscher, P., Clark-Lewis, I. (2001) Agonistic and antagonistic activities
of chemokines. J. Leukoc. Biol. 69, 881– 884.
11. Zhang, K., McQuibban, G. A., Silva, C., Butler, G. S., Johnston, J. B.,
Holden, J., Clark-Lewis, I., Overall, C. M., Power, C. (2003) HIV-induced
metalloproteinase processing of the chemokine stromal cell derived fac-
tor-1 causes neurodegeneration. Nat. Neurosci. 6, 1064 –1071.
12. McQuibban, G. A., Butler, G. S., Gong, J. H., Bendall, L., Power, C.,
Clark-Lewis, I., Overall, C. M. (2001) Matrix metalloproteinase activity
inactivates the CXC chemokine stromal cell-derived factor-1. J. Biol.
Chem. 276, 43503– 43508.
13. Dean, R. A., Overall, C. M. (2007) Proteomics discovery of metallopro-
teinase substrates in the cellular context by iTRAQTM labeling reveals a
diverse MMP-2 substrate degradome. Mol. Cell. Proteomics 6, 611– 623.
14. Van den Steen, P. E., Proost, P., Wuyts, A., Van Damme, J., Opdenakker,
G. (2000) Neutrophil gelatinase B potentiates interleukin-8 tenfold by
aminoterminal processing, whereas it degrades CTAP-III, PF-4, and
GRO-and leaves RANTES and MCP-2 intact. Blood 96, 2673–2681.
15. Van Den Steen, P. E., Wuyts, A., Husson, S. J., Proost, P., Van Damme,
J., Opdenakker, G. (2003) Gelatinase B/MMP-9 and neutrophil collage-
nase/MMP-8 process the chemokines human GCP-2/CXCL6, ENA-78/
CXCL5 and mouse GCP-2/LIX and modulate their physiological activities.
Eur. J. Biochem. 270, 3739 –3749.
16. Van den Steen, P. E., Husson, S. J., Proost, P., Van Damme, J., Opde-
nakker, G. (2003) Carboxyterminal cleavage of the chemokines MIG and
Van Lint and Libert Matrix metalloproteinases in inflammation 5
IP-10 by gelatinase B and neutrophil collagenase. Biochem. Biophys. Res.
Commun. 310, 889 896.
17. Liao, F., Rabin, R. L., Yannelli, J. R., Koniaris, L. G., Vanguri, P., Farber,
J. M. (1995) Human Mig chemokine: biochemical and functional charac-
terization. J. Exp. Med. 182, 1301–1314.
18. McQuibban, G. A., Gong, J. H., Tam, E. M., McCulloch, C. A., Clark-
Lewis, I., Overall, C. M. (2000) Inflammation dampened by gelatinase A
cleavage of monocyte chemoattractant protein-3. Science 289, 1202–
1206.
19. McQuibban, G. A., Gong, J. H., Wong, J. P., Wallace, J. L., Clark-Lewis,
I., Overall, C. M. (2002) Matrix metalloproteinase processing of monocyte
chemoattractant proteins generates CC chemokine receptor antagonists
with anti-inflammatory properties in vivo. Blood 100, 1160 –1167.
20. Tam, E. M., Morrison, C. J., Wu, Y. I., Stack, M. S., Overall, C. M. (2004)
Membrane protease proteomics: isotope-coded affinity tag MS identifica-
tion of undescribed MT1-matrix metalloproteinase substrates. Proc. Natl.
Acad. Sci. USA 101, 6917– 6922.
21. Tester, A. M., Cox, J. H., Connor, A. R., Starr, A. E., Dean, R. A., Puente,
X. S., Lopez-Otin, C., Overall, C. M. (2007) LPS responsiveness and
neutrophil chemotaxis in vivo require PMN MMP-8 activity. PLoS ONE 2,
e312.
22. Balbin, M., Fueyo, A., Tester, A. M., Pendas, A. M., Pitiot, A. S.,
Astudillo, A., Overall, C. M., Shapiro, S. D., Lopez-Otin, C. (2003) Loss of
collagenase-2 confers increased skin tumor susceptibility to male mice.
Nat. Genet. 35, 252–257.
23. Goger, B., Halden, Y., Rek, A., Mosl, R., Pye, D., Gallagher, J., Kungl,
A. J. (2002) Different affinities of glycosaminoglycan oligosaccharides for
monomeric and dimeric interleukin-8: a model for chemokine regulation at
inflammatory sites. Biochemistry 41, 1640 –1646.
24. Proudfoot, A. E. (2006) The biological relevance of chemokine-proteogly-
can interactions. Biochem. Soc. Trans. 34, 422– 426.
25. Li, Q., Park, P. W., Wilson, C. L., Parks, W. C. (2002) Matrilysin shedding
of syndecan-1 regulates chemokine mobilization and transepithelial efflux
of neutrophils in acute lung injury. Cell 111, 635– 646.
26. Fitzgerald, M. L., Wang, Z., Park, P. W., Murphy, G., Bernfield, M. (2000)
Shedding of syndecan-1 and -4 ectodomains is regulated by multiple
signaling pathways and mediated by a TIMP-3-sensitive metalloprotein-
ase. J. Cell Biol. 148, 811– 824.
27. Yu, W. H., Woessner Jr., J. F. (2000) Heparan sulfate proteoglycans as
extracellular docking molecules for matrilysin (matrix metalloproteinase
7). J. Biol. Chem. 275, 4183– 4191.
28. Brule, S., Charnaux, N., Sutton, A., Ledoux, D., Chaigneau, T., Saffar, L.,
Gattegno, L. (2006) The shedding of syndecan-4 and syndecan-1 from
HeLa cells and human primary macrophages is accelerated by SDF-1/
CXCL12 and mediated by the matrix metalloproteinase-9. Glycobiology
16, 488 –501.
29. Endo, K., Takino, T., Miyamori, H., Kinsen, H., Yoshizaki, T., Furukawa,
M., Sato, H. (2003) Cleavage of syndecan-1 by membrane type matrix
metalloproteinase-1 stimulates cell migration. J. Biol. Chem. 278,
40764 40770.
30. Van Lint, P., Wielockx, B., Puimege, L., Noel, A., Lopez-Otin, C., Libert,
C. (2005) Resistance of collagenase-2 (matrix metalloproteinase-8)-defi-
cient mice to TNF-induced lethal hepatitis. J. Immunol. 175, 7642–
7649.
31. Corry, D. B., Rishi, K., Kanellis, J., Kiss, A., Song Lz, L. Z., Xu, J., Feng,
L., Werb, Z., Kheradmand, F. (2002) Decreased allergic lung inflamma-
tory cell egression and increased susceptibility to asphyxiation in MMP2-
deficiency. Nat. Immunol. 3, 347–353.
32. Corry, D. B., Kiss, A., Song, L. Z., Song, L., Xu, J., Lee, S. H., Werb, Z.,
Kheradmand, F. (2004) Overlapping and independent contributions of
MMP2 and MMP9 to lung allergic inflammatory cell egression through
decreased CC chemokines. FASEB J. 18, 995–997.
33. Stefanidakis, M., Ruohtula, T., Borregaard, N., Gahmberg, C. G., Koi-
vunen, E. (2004) Intracellular and cell surface localization of a complex
between M2 integrin and promatrix metalloproteinase-9 progelatinase
in neutrophils. J. Immunol. 172, 7060 –7068.
34. Vermaelen, K. Y., Cataldo, D., Tournoy, K., Maes, T., Dhulst, A., Louis,
R., Foidart, J. M., Noel, A., Pauwels, R. (2003) Matrix metalloproteinase-
9-mediated dendritic cell recruitment into the airways is a critical step in
a mouse model of asthma. J. Immunol. 171, 1016 –1022.
35. Cataldo, D. D., Tournoy, K. G., Vermaelen, K., Munaut, C., Foidart, J. M.,
Louis, R., Noel, A., Pauwels, R. A. (2002) Matrix metalloproteinase-9
deficiency impairs cellular infiltration and bronchial hyperresponsiveness
during allergen-induced airway inflammation. Am. J. Pathol. 161, 491–
498.
36. Haro, H., Crawford, H. C., Fingleton, B., MacDougall, J. R., Shinomiya,
K., Spengler, D. M., Matrisian, L. M. (2000) Matrix metalloproteinase-3-
dependent generation of a macrophage chemoattractant in a model of
herniated disc resorption. J. Clin. Invest. 105, 133–141.
37. Warner, R. L., Beltran, L., Younkin, E. M., Lewis, C. S., Weiss, S. J.,
Varani, J., Johnson, K. J. (2001) Role of stromelysin 1 and gelatinase B in
experimental acute lung injury. Am. J. Respir. Cell Mol. Biol. 24, 537–
544.
38. Hautamaki, R. D., Kobayashi, D. K., Senior, R. M., Shapiro, S. D. (1997)
Requirement for macrophage elastase for cigarette smoke-induced emphy-
sema in mice. Science 277, 2002–2004.
39. Houghton, A. M., Quintero, P. A., Perkins, D. L., Kobayashi, D. K.,
Kelley, D. G., Marconcini, L. A., Mecham, R. P., Senior, R. M., Shapiro,
S. D. (2006) Elastin fragments drive disease progression in a murine
model of emphysema. J. Clin. Invest. 116, 753–759.
40. Hunninghake, G. W., Davidson, J. M., Rennard, S., Szapiel, S., Gadek,
J. E., Crystal, R. G. (1981) Elastin fragments attract macrophage precur-
sors to diseased sites in pulmonary emphysema. Science 212, 925–927.
41. Senior, R. M., Griffin, G. L., Mecham, R. P. (1980) Chemotactic activity
of elastin-derived peptides. J. Clin. Invest. 66, 859 862.
42. Weathington, N. M., van Houwelingen, A. H., Noerager, B. D., Jackson,
P. L., Kraneveld, A. D., Galin, F. S., Folkerts, G., Nijkamp, F. P., Blalock,
J. E. (2006) A novel peptide CXCR ligand derived from extracellular
matrix degradation during airway inflammation. Nat. Med. 12, 317–323.
43. Adair-Kirk, T. L., Atkinson, J. J., Broekelmann, T. J., Doi, M., Tryggvason,
K., Miner, J. H., Mecham, R. P., Senior, R. M. (2003) A site on laminin
5, AQARSAASKVKVSMKF, induces inflammatory cell production of
matrix metalloproteinase-9 and chemotaxis. J. Immunol. 171, 398 406.
44. Clark, R. A., Wikner, N. E., Doherty, D. E., Norris, D. A. (1988) Cryptic
chemotactic activity of fibronectin for human monocytes resides in the
120-kDa fibroblastic cell-binding fragment. J. Biol. Chem. 263, 12115–
12123.
45. Nakagawa, H., Takano, K., Kuzumaki, H. (1999) A 16-kDa fragment of
collagen type XIV is a novel neutrophil chemotactic factor purified from
rat granulation tissue. Biochem. Biophys. Res. Commun. 256, 642– 645.
46. Postlethwaite, A. E., Kang, A. H. (1976) Collagen- and collagen peptide-
induced chemotaxis of human blood monocytes. J. Exp. Med. 143,
1299 –1307.
47. Banda, M. J., Rice, A. G., Griffin, G. L., Senior, R. M. (1988) 1-Pro-
teinase inhibitor is a neutrophil chemoattractant after proteolytic inacti-
vation by macrophage elastase. J. Biol. Chem. 263, 4481– 4484.
48. Pruijt, J. F., Fibbe, W. E., Laterveer, L., Pieters, R. A., Lindley, I. J.,
Paemen, L., Masure, S., Willemze, R., Opdenakker, G. (1999) Prevention
of interleukin-8-induced mobilization of hematopoietic progenitor cells in
rhesus monkeys by inhibitory antibodies against the metalloproteinase
gelatinase B (MMP-9). Proc. Natl. Acad. Sci. USA 96, 10863–10868.
49. Pruijt, J. F., Verzaal, P., van Os, R., de Kruijf, E. J., van Schie, M. L.,
Mantovani, A., Vecchi, A., Lindley, I. J., Willemze, R., Starckx, S.,
Opdenakker, G., Fibbe, W. E. (2002) Neutrophils are indispensable for
hematopoietic stem cell mobilization induced by interleukin-8 in mice.
Proc. Natl. Acad. Sci. USA 99, 6228 6233.
50. Heissig, B., Hattori, K., Dias, S., Friedrich, M., Ferris, B., Hackett, N. R.,
Crystal, R. G., Besmer, P., Lyden, D., Moore, M. A., Werb, Z., Rafii, S.
(2002) Recruitment of stem and progenitor cells from the bone marrow
niche requires MMP-9 mediated release of kit-ligand. Cell 109, 625–
637.
51. Clutter, S. D., Fortney, J., Gibson, L. F. (2005) MMP-2 is required for bone
marrow stromal cell support of pro-B-cell chemotaxis. Exp. Hematol. 33,
1192–1200.
52. Agrawal, S., Anderson, P., Durbeej, M., van Rooijen, N., Ivars, F.,
Opdenakker, G., Sorokin, L. M. (2006) Dystroglycan is selectively cleaved
at the parenchymal basement membrane at sites of leukocyte extravasation
in experimental autoimmune encephalomyelitis. J. Exp. Med. 203, 1007–
1019.
53. Moss, M. L., Jin, S. L., Milla, M. E., Bickett, D. M., Burkhart, W., Carter,
H. L., Chen, W. J., Clay, W. C., Didsbury, J. R., Hassler, D., et al. (1997)
Cloning of a disintegrin metalloproteinase that processes precursor tumor-
necrosis factor-.Nature 385, 733–736.
54. Ruuls, S. R., Hoek, R. M., Ngo, V. N., McNeil, T., Lucian, L. A.,
Janatpour, M. J., Korner, H., Scheerens, H., Hessel, E. M., Cyster, J. G.,
McEvoy, L. M., Sedgwick, J. D. (2001) Membrane-bound TNF supports
secondary lymphoid organ structure but is subservient to secreted TNF in
driving autoimmune inflammation. Immunity 15, 533–543.
55. Mohan, M. J., Seaton, T., Mitchell, J., Howe, A., Blackburn, K., Burkhart,
W., Moyer, M., Patel, I., Waitt, G. M., Becherer, J. D., Moss, M. L., Milla,
M. E. (2002) The tumor necrosis factor-converting enzyme (TACE): a
unique metalloproteinase with highly defined substrate selectivity. Bio-
chemistry 41, 9462–9469.
56. Haro, H., Crawford, H. C., Fingleton, B., Shinomiya, K., Spengler, D. M.,
Matrisian, L. M. (2000) Matrix metalloproteinase-7-dependent release of
6 Journal of Leukocyte Biology Volume 82, December 2007 http://www.jleukbio.org
tumor necrosis factor-in a model of herniated disc resorption. J. Clin.
Invest. 105, 143–150.
57. Gearing, A. J., Beckett, P., Christodoulou, M., Churchill, M., Clements, J.,
Davidson, A. H., Drummond, A. H., Galloway, W. A., Gilbert, R., Gordon,
J. L., et al. (1994) Processing of tumor necrosis factor-precursor by
metalloproteinases. Nature 370, 555–557.
58. English, W. R., Puente, X. S., Freije, J. M., Knauper, V., Amour, A.,
Merryweather, A., Lopez-Otin, C., Murphy, G. (2000) Membrane type 4
matrix metalloproteinase (MMP17) has tumor necrosis factor-convertase
activity but does not activate pro-MMP2. J. Biol. Chem. 275, 14046
14055.
59. Fantuzzi, G., Ku, G., Harding, M. W., Livingston, D. J., Sipe, J. D., Kuida,
K., Flavell, R. A., Dinarello, C. A. (1997) Response to local inflammation
of IL-1 -converting enzyme-deficient mice. J. Immunol. 158, 1818
1824.
60. Nylander-Lundqvist, E., Back, O., Egelrud, T. (1996) IL-1 activation in
human epidermis. J. Immunol. 157, 1699 –1704.
61. Schonbeck, U., Mach, F., Libby, P. (1998) Generation of biologically
active IL-1 by matrix metalloproteinases: a novel caspase-1-indepen-
dent pathway of IL-1 processing. J. Immunol. 161, 3340 –3346.
62. Ito, A., Mukaiyama, A., Itoh, Y., Nagase, H., Thogersen, I. B., Enghild,
J. J., Sasaguri, Y., Mori, Y. (1996) Degradation of interleukin 1by matrix
metalloproteinases. J. Biol. Chem. 271, 14657–14660.
63. Sheppard, D. (2006) Transforming growth factor : a central modulator of
pulmonary and airway inflammation and fibrosis. Proc. Am. Thorac. Soc.
3, 413– 417.
64. Yu, Q., Stamenkovic, I. (2000) Cell surface-localized matrix metallopro-
teinase-9 proteolytically activates TGF-and promotes tumor invasion
and angiogenesis. Genes Dev. 14, 163–176.
65. Maeda, S., Dean, D. D., Gomez, R., Schwartz, Z., Boyan, B. D. (2002) The
first stage of transforming growth factor 1 activation is release of the large
latent complex from the extracellular matrix of growth plate chondrocytes
by matrix vesicle stromelysin-1 (MMP-3). Calcif. Tissue Int. 70, 54 65.
66. Karsdal, M. A., Larsen, L., Engsig, M. T., Lou, H., Ferreras, M., Lochter,
A., Delaisse, J. M., Foged, N. T. (2002) Matrix metalloproteinase-depen-
dent activation of latent transforming growth factor-controls the conver-
sion of osteoblasts into osteocytes by blocking osteoblast apoptosis. J. Biol.
Chem. 277, 44061– 44067.
67. Dallas, S. L., Rosser, J. L., Mundy, G. R., Bonewald, L. F. (2002)
Proteolysis of latent transforming growth factor-(TGF-)-binding pro-
tein-1 by osteoclasts. A cellular mechanism for release of TGF-from
bone matrix. J. Biol. Chem. 277, 21352–21360.
68. Imai, K., Hiramatsu, A., Fukushima, D., Pierschbacher, M. D., Okada, Y.
(1997) Degradation of decorin by matrix metalloproteinases: identification
of the cleavage sites, kinetic analyses and transforming growth factor-1
release. Biochem. J. 322, 809 814.
69. Velasco-Loyden, G., Arribas, J., Lopez-Casillas, F. (2004) The shedding of
glycan is regulated by pervanadate and mediated by membrane type
matrix metalloprotease-1. J. Biol. Chem. 279, 7721–7733.
70. Lopez-Casillas, F., Payne, H. M., Andres, J. L., Massague, J. (1994)
Glycan can act as a dual modulator of TGF-access to signaling
receptors: mapping of ligand binding and GAG attachment sites. J. Cell
Biol. 124, 557–568.
71. Nelissen, I., Martens, E., Van den Steen, P. E., Proost, P., Ronsse, I.,
Opdenakker, G. (2003) Gelatinase B/matrix metalloproteinase-9 cleaves
interferon-and is a target for immunotherapy. Brain 126, 1371–1381.
72. Orlando, S., Sironi, M., Bianchi, G., Drummond, A. H., Boraschi, D.,
Yabes, D., Mantovani, A. (1997) Role of metalloproteases in the release of
the IL-1 type II decoy receptor. J. Biol. Chem. 272, 31764 –31769.
73. Bellehumeur, C., Collette, T., Maheux, R., Mailloux, J., Villeneuve, M.,
Akoum, A. (2005) Increased soluble interleukin-1 receptor type II prote-
olysis in the endometrium of women with endometriosis. Hum. Reprod.
20, 1177–1184.
74. Sheu, B. C., Hsu, S. M., Ho, H. N., Lien, H. C., Huang, S. C., Lin, R. H.
(2001) A novel role of metalloproteinase in cancer-mediated immunosup-
pression. Cancer Res. 61, 237–242.
75. Mortier, E., Bernard, J., Plet, A., Jacques, Y. (2004) Natural, proteolytic
release of a soluble form of human IL-15 receptor -chain that behaves as
a specific, high affinity IL-15 antagonist. J. Immunol. 173, 1681–1688.
Van Lint and Libert Matrix metalloproteinases in inflammation 7
... Наиболее часто при инсульте имеет значение желатиназа (коллагеназа IV) -ММП-9. Она участвует в разрушении ГЭБ, повышает проницаемость его с развитием отека мозговой ткани, эксайтоксичность, оксидативный стресс, нарушение репаративного действия ДНК и часто лейкоцитарную инфильтрацию [4,6,12]. Проницаемость ГЭБ и отек мозга часто способствует появлению кровоизлияний [5]. ...
... И чем больше идет площадь поражения тканевого матрикса или выраженный цереброваскулярный процесс, тем возможно происходит больше продукции ММП-9. ММП-9 повышает проницаемость ГЭБ, вызывает отек тканей, эксайтотоксичность и оксидативный стресс [4,6,12]. Происходит дальнейшее разрушение внеклеточного матрикса с образованием порочного круга. ...
Article
Было обследовано 87 больных с цереброваскулярными заболеваниями и инсультами в возрасте от 50 до 78 лет, средний возраст 67,9±4,8 лет. Выявленная тенденция к повышению уровня ММП-9, ТИМП-1 и их соотношения в сыворотке у больных с цереброваскулярными заболеваниями можно расценивать как предиктор выраженности сосудисто-мозгового процесса.
... They play a role in a number of physiological processes: tissue remodeling, reproduction, tissue resorption, angiogenesis, proliferation, migration and differentiation of cells, apoptosis, inhibition of tumor growth. Involved in the cleavage of membrane receptors, the release of apoptotic ligands such as FAS, as well as the activation and deactivation of chemokines and cytokines [33]. Multidirectional changes in MMP activity are observed in various pathological processes, incl. ...
Article
Цель. На основании оценки экспрессии основных коллагенов и матричных металлопротеиназ установить эффективность экспериментального использования при лечении хронических ран промежности локальной клеточной терапии и высокоинтенсивного лазерного излучения определенных параметров. Материалы и методы. После моделирования хронической раны промежности у животных первой группы А (n=24) на область хронической раны осуществляли воздействие расфокусированным лазерным излучением длиной волны 1560 нм (0,1 с / 0,1 с, 5 Вт, диаметр светового пятна 0,7 см, дистанция 2 см, суммарная экспозиция 10 с). Во второй группе животных Б (n=24) в края и дно раны инъекционно вводили суспензию аллогенных мезенхимальных стволовых клеток жировой ткани (МСК ЖТ) (1 мл, 500 тыс/мл). В третьей группе животных В (n=24) в края и дно раны вводили 2 мл аллогенной обогащенной тромбоцитами и лейкоцитами плазмы (L-PRP). В четвертой группе Г (n=24) в края раны однократно вводили 2 мл физиологического раствора поваренной соли (плацебо-контроль). В различные сроки после воздействия проведена оценка молекулярно-биологических маркеров в тканях лабораторных животных иммуногистохимическим методом с использованием моно- и поликлональных антител. Результаты. Иммуногистохимическая оценка тканей с оценкой уровня экспрессии коллагенов I и III типа в различные сроки выявила его разнонаправленный (разноуровневый) характер, зависящий от вида применяемого лечебного метода, сроков проявления биологического эффекта и специфики индивидуальной реакции лабораторных животных на лечебное воздействие. При этом использование всех лечебных технологий в динамике достоверно приводит к снижению активности продукции ММР-1 и ММР-9 в сравнении с группой плацебо-контроля, коррелируя с изменениями уровня экспрессии коллагенов, инволюцией воспалительно-дегенеративных процессов в ране и уровнем микробной колонизации ран. В большей степени этот процесс наблюдается в группе животных с использованием лазерного излучения (во все сроки, начиная с 5-х суток наблюдения). Заключение. Ввиду выявленного в экспериментальных исследованиях значимого регенераторного эффекта в отношении хронических ран промежности анализируемых методов следует считать целесообразным местное использование в клинических условиях лазерного излучения и клеточной терапии. Purpose. Based on the assessment of the expression of the main collagens and matrix metalloproteinases, to establish the effectiveness of the experimental use of local cell therapy and high-intensity laser radiation of certain parameters in the treatment of chronic perineal wounds. Materials and methods. After modeling a chronic wound of the perineum in animals of the first group A (n=24), the area of the chronic wound was exposed to defocused laser radiation with a wavelength of 1560 nm (0.1 s / 0.1 s, 5 W, light spot diameter 0.7 cm, distance 2 cm, total exposure 10 s). In the second group of animals B (n=24), a suspension of allogeneic adipose tissue mesenchymal stem cells (ATMSC) (1 ml, 500 thousand/ml) was injected into the edges and bottom of the wound. In the third group of animals B (n=24), 2 ml of allogeneic platelet- and leukocyte-rich plasma (L-PRP) was injected into the edges and bottom of the wound. In the fourth group G (n=24), 2 ml of physiological sodium chloride solution was injected once into the edges of the wound (placebo control). At various times after exposure, molecular biological markers in the tissues of laboratory animals were assessed using the immunohistochemical method using mono- and polyclonal antibodies. Results. Immunohistochemical assessment of tissues with assessment of the level of expression of type I and III collagens at different times has established its multidirectional (multi-level) nature, depending on the type of therapeutic method used, the timing of the manifestation of the biological effect and the specifics of the individual reaction of laboratory animals to the therapeutic effect. At the same time, the use of all therapeutic technologies in dynamics reliably leads to a decrease in the activity of MMP-1 and MMP-9 production in comparison with the placebo control group, correlating with changes in the level of collagen expression, involution of inflammatory-degenerative processes in the wound and the level of microbial colonization of wounds. To a greater extent, this process is observed in the group of animals using laser radiation (at all times, starting from the 5th day of observation). Conclusion. In view of the significant regenerative effect of the analyzed methods in relation to chronic perineal wounds identified in experimental studies, local use of laser radiation and cell therapy in clinical conditions should be considered appropriate.
... Collectively, these enzymes can degrade a wide range of extracellular matrix proteins as well as a variety of bioactive molecules. They have been linked to the cleavage of cell surface receptors, the release of apoptotic ligands (such as the FAS ligand), and the inactivation of chemokines and cytokines [19] [20]. Additionally, it is believed that MMPs are important for cellular behaviors such angiogenesis, apoptosis, differentiation, migration (adhesion/dispersion), and host defence [21] [22]. ...
... Additional evidence for immunomodulation of C. angustifolia is increasing in the fold expression of MMP family, PECAM-1, TNF alpha, VEGFR2 and µPAR. These inflammations proteins have lot of important and crucial role in the immune response and defense, through play role for recruitment of leukocyte, cytokine production, inflammation process, adhesion of platelet (33,34). The macrophage plasminogen activators is µPAR (urokineas) that control for phagocytosis to the chronic inflammation leads to immunomodulation and development of immune response by supportive matrix generation and control to the cells migration, adhesion, proliferation the process in the inflammatory response (35). ...
Article
Full-text available
Cassia angustifolia Vahl. is a medicinal plant known for its efficacy in treating various, including respiratory conditions and skin inflammation. It possesses antibacterial and anticancer properties. This work investigated the immunomodulatory and anti-inflammatory effects of C. angustifolia. The ethanol leaf extract of C. angustifolia was utilized to examine gene expression related to angiogenesis cytokines in RAW 264.7 macrophage cells. The results demonstrated a significant increase in the viability of treated macrophage RAW 264.7 cells, accompanied by an improvement in angiogenesis cytokines expression and a dose-dependent inhibition of nitric oxide production. GC-MS analysis identified 11 active components within the extract, each exhibiting distinct biological activities such as antioxidant, antitumor and anti-inflammatory effects. Notable compounds include hexadecanoic acid, 2-pentadecanone, phthalic acid, oxalic acid, carbonic acid, tricosane, undecanal and many others. In conclusion, ethanol leaf extract of C. angustifolia exhibits immunomodulatory and anti-inflammatory effects by inhibiting nitric oxide production and enhancing the expression of angiogenesis cytokines.
Article
Full-text available
Funduscopic diseases, including diabetic retinopathy (DR) and age‐related macular degeneration (AMD), significantly impact global visual health, leading to impaired vision and irreversible blindness. Delivering drugs to the posterior segment of the eye remains a challenge due to the presence of multiple physiological and anatomical barriers. Conventional drug delivery methods often prove ineffective and may cause side effects. Nanomaterials, characterized by their small size, large surface area, tunable properties, and biocompatibility, enhance the permeability, stability, and targeting of drugs. Ocular nanomaterials encompass a wide range, including lipid nanomaterials, polymer nanomaterials, metal nanomaterials, carbon nanomaterials, quantum dot nanomaterials, and so on. These innovative materials, often combined with hydrogels and exosomes, are engineered to address multiple mechanisms, including macrophage polarization, reactive oxygen species (ROS) scavenging, and anti‐vascular endothelial growth factor (VEGF). Compared to conventional modalities, nanomedicines achieve regulated and sustained delivery, reduced administration frequency, prolonged drug action, and minimized side effects. This study delves into the obstacles encountered in drug delivery to the posterior segment and highlights the progress facilitated by nanomedicine. Prospectively, these findings pave the way for next‐generation ocular drug delivery systems and deeper clinical research, aiming to refine treatments, alleviate the burden on patients, and ultimately improve visual health globally.
Article
Статистика передчасного народження дітей відрізняється в різних країнах і в різні періоди часу, проте це залишається серйозною медичною проблемою. За даними Всесвітньої організації охорони здоров'я, близько 10% вагітностей завершуються передчасним народженням у всьому світі. Це означає, що більше 1 з 10 дітей народжується передчасно. Передчасним народженням вважається те, яке відбувається до 37-го тижня вагітності. Проте є і більш ранні випадки, коли дитина народжується з недорозвиненістю органів та систем. На передчасні пологи припадає більшість неонатальної захворюваності та смертності в розвинених країнах. У статті досліджується фізіологічні механізми впливу широковідомих гуморальних факторів пологів таких як окситоцин, простагландини, естрогени, прогестерон, релаксин, кортикотропін-релізинг гормон та ендотеліальні фактори як оксид азоту і ендотелін. Окрім того розглядається участь менш досліджених факторів регулюючих пологову діяльність таких як 15-гідроксипростагландин дегідрогеназа, хоріонічний гонадотропін людини, активний білок сурфактанта, активатор тромбоцитів плода, каспази. Проаналізовані фактори розширюють розуміння багатокомпонентності індукторів пологової діяльності та складності гуморальної взаємодії між плодом та організмом породілля. Описанні у статті процеси гуморальної регуляції періодів пологової діяльності можуть потенційно використовуватися для протидії передчасним пологам.
Article
Objectives To evaluate the differences in vaginal matrix metalloproteinases (MMP) and tissue inhibitors of metalloproteinases (TIMPs) in pregnant patients with a history of prior preterm birth compared with controls. Methods A prospective cohort pilot study recruited patients during prenatal care with history of prior spontaneous preterm birth (high-risk group) or no history of preterm birth (low-risk/controls). Inclusion criteria were singleton gestation at 11–16 weeks and between 18 and 55 years of age. Exclusion criteria were diabetes mellitus, hypertension, diseases affecting the immune response or acute vaginitis. A vaginal wash was performed at time of enrollment, and patients were followed through delivery. Samples were analyzed using semi-quantitative analysis of MMPS and TIMPS. The study was approved by the IRB and a p value <0.05 was considered significant. Results A total of 48 pregnant patients were recruited: 16 with a history of preterm birth (high-risk group) and 32 with no history of preterm birth (low-risk group/controls). Groups were similar in age, race, BMI, and delivery mode. The high-risk group had more multiparous women (100 vs. 68.8 %; p=0.02), a greater preterm birth rate (31.2 vs. 6.3 %; p=0.02), and a lower birth weight (2,885 ± 898 g vs. 3,480 ± 473 g; p=0.02). Levels of vaginal MMP-9 were greater in high-risk patients than low-risk patients (74.9 % ± 27.0 vs. 49.4 % ± 31.1; p=0.01). When dividing the cohort into patients that had a spontaneous preterm birth (7/48, 14.6 %) vs. those with a term delivery (41/48, 85.4 %), the vaginal MMP-9 remained elevated in the cohort that experienced a preterm birth (85.46 %+19.79 vs. 53.20 %+31.47; p=0.01). There were no differences in the other MMPS and in TIMPs between high and low-risk groups. Conclusions There was an increase in vaginal MMP-9 during early pregnancy in those at high risk for preterm birth and in those who delivered preterm, regardless of prior pregnancy outcome. Vaginal MMP-9 may have potential as a marker of increased risk of preterm birth.
Article
IL-1 beta is produced as an inactive 31-kDa precursor processed to active 18-kDa IL-1 beta by proteolytic cleavage, catalyzed by the highly specific IL-1-converting enzyme (ICE). In vitro activation of IL-1 beta can also be obtained by other proteases. Human keratinocytes express IL-1 beta, but not active ICE. The role played by IL-1 beta produced by keratinocytes has therefore been unclear. We asked whether normal human plantar stratum corneum contains biologically active IL-1 beta and, if so, by which mechanism this IL-1 beta is activated. Crude extracts and partially purified preparations from which IL-1 alpha had been removed were used. Biologic IL-1 activity was measured as the ability to induce expression of E-selectin in HUVEC. The crude extract contained IL-1-like activity that could be partially abolished with Abs to IL-1 alpha or IL-1 beta and totally inhibited with a mixture of the two Abs. IL-1 activity in the partially purified preparation was totally inhibited by Abs to IL-1 beta. The specific IL-1 beta activity in the two preparations was 60 to 85% of the sp. act. of recombinant human IL-1 beta. IL-1 beta from plantar stratum corneum had a slightly higher molecular mass than recombinant mature IL-1 beta. Its isoelectric point was approximately 6.1 compared with 6.9 for rIL-1 beta. We conclude that human plantar stratum corneum contains biologically active IL-1 beta that has been activated by an alternative mechanism that does not involve ICE. This is, to our knowledge, the first report of an alternative mechanism of IL-1 beta activation occurring in vivo.
Article
Chemokines are mediators in inflammatory and autoimmune disorders. Aminoterminal truncation of chemokines results in altered specific activities and receptor recognition patterns. Truncated forms of the CXC chemokine interleukin (IL)-8 are more active than full-length IL-8 (1-77), provided the Glu-Leu-Arg (ELR) motif remains intact. Here, a positive feedback loop is demonstrated between gelatinase B, a major secreted matrix metalloproteinase (MMP-9) from neutrophils, and IL-8, the prototype chemokine active on neutrophils. Natural human neutrophil progelatinase B was purified to homogeneity and activated by stromelysin-1. Gelatinase B truncated IL-8(1-77) into IL-8(7-77), resulting in a 10- to 27-fold higher potency in neutrophil activation, as measured by the increase in intracellular Ca++concentration, secretion of gelatinase B, and neutrophil chemotaxis. This potentiation correlated with enhanced binding to neutrophils and increased signaling through CXC chemokine receptor-1 (CXCR1), but it was significantly less pronounced on a CXCR2-expressing cell line. Three other CXC chemokines—connective tissue-activating peptide-III (CTAP-III), platelet factor-4 (PF-4), and GRO-α—were degraded by gelatinase B. In contrast, the CC chemokines RANTES and monocyte chemotactic protein-2 (MCP-2) were not digested by this enzyme. The observation of differing effects of neutrophil gelatinase B on the proteolysis of IL-8 versus other CXC chemokines and on CXC receptor usage by processed IL-8 yielded insights into the relative activities of chemokines. This led to a better understanding of regulator (IL-8) and effector molecules (gelatinase B) of neutrophils and of mechanisms underlying leukocytosis, shock syndromes, and stem cell mobilization by IL-8.
Article
Monocyte chemoattractant protein (MCP)–3 is inactivated upon cleavage by the matrix metalloproteinase (MMP) gelatinase A (MMP-2). We investigated the susceptibility to proteolytic processing of the 4 human MCPs by 8 recombinant MMPs to determine whether MCP-3 is an isolated example or represents a general susceptibility of chemokines to proteolytic inactivation by these important inflammatory proteases. In addition to MMP-2, MCP-3 is efficiently cleaved by membrane type 1 (MT1)–MMP, the cellular activator of MMP-2, and by collagenase-1 and collagenase-3 (MMP-1, MMP-13) and stromelysin-1 (MMP-3). Specificity was shown by absence of cleavage by matrilysin (MMP-7) and the leukocytic MMPs neutrophil collagenase (MMP-8) and gelatinase B (MMP-9). The closely related chemokines MCP-1, MCP-2, and MCP-4 were not cleaved by MMP-2 or MT1-MMP, but were cleaved by MMP-1 and MMP-3 with varying efficiency. MCPs were typically cleaved between residues 4 and 5, but MCP-4 was further processed at Val7-Pro8. Synthetic MCP analogs corresponding to the MMP-cleaved forms bound CC chemokine receptor (CCR)–2 and CCR-3, but lacked chemoattractant activity in pre-B cells transfected with CCR-2 and CCR-3 or in THP-1 monocytic cells, a transformed leukemic cell line. Moreover, the truncated products of MCP-2 and MCP-4, like MCP-3, were potent antagonists of their cognate CC chemokine receptors in transwell cell migration assays in vitro. When they were injected 24 hours after the initiation of carrageenan-induced inflammation in rat paws, their in vivo antagonist activities were revealed by a greater than 66% reduction in inflammatory edema progression after 12 hours. We propose that MMPs have an important role in modulating inflammatory and immune responses by processing chemokines in wound healing and in disease.
Article
MTl-MMP-deficient mice were found to have severe defects in skeletal development and angiogenesis. The craniofacial, appendicular skeletons were severely affected, leading to a short and domed skull, marked deceleration of postnatal growth and death by 3 weeks of age. Shortening of bones is a consequence of decreased chondrocytes proliferation in the proliferative zone of growth plates. Defective vascular invasion of cartilage leads to enlargement of hypertrophie zone in growth plates and delayed formation of secondary ossification centers in long bones. In an in vivo corneal assay, null mice did not have angiogenic response to implanted FGF-b, suggesting that the defect in angiogenesis is not restricted to cartilage. In tissues from null mice, activation of proMMP-2 was deficient, suggesting MT1-MMP is essential for proMMP-2 activation.
Article
Dendritic cells (DCs) appear to be strategically implicated in allergic diseases, including asthma. Matrix metalloproteinase (MMP)-9 mediates transmigration of inflammatory leukocytes across basement membranes. This study investigated the role of MMP-9 in airway DC trafficking during allergen-induced airway inflammation. MMP-9 gene deletion affected the trafficking of pulmonary DCs in a specific way: only the inflammatory transmigration of DCs into the airway lumen was impaired, whereas DC-mediated transport of airway Ag to the thoracic lymph nodes remained unaffected. In parallel, the local production of the Th2-attracting chemokine CC chemokine ligand 17/thymus and activation-regulated chemokine, which was highly concentrated in purified lung DCs, fell short in the airways of allergen-exposed MMP-9(-/-) mice. This was accompanied by markedly reduced peribronchial eosinophilic infiltrates and impaired allergen-specific IgE production. We conclude that the specific absence of MMP-9 activity inhibits the development of allergic airway inflammation by impairing the recruitment of DCs into the airways and the local production of DC-derived proallergic chemokines.
Article
Betaglycan, also known as the TGF-beta type III receptor, is a membrane- anchored proteoglycan that presents TGF-beta to the type II signaling receptor, a transmembrane serine/threonine kinase. The betaglycan extracellular region, which can be shed by cells into the medium, contains a NH2-terminal domain related to endoglin and a COOH-terminal domain related to uromodulin, sperm receptors Zp2 and 3, and pancreatic secretory granule GP-2 protein. We identified residues Ser535 and Ser546 in the uromodulin-related region as the glycosaminoglycan (GAG) attachment sites. Their mutation to alanine prevents GAG attachment but does not interfere with betaglycan stability or ability to bind and present TGF-beta to receptor II. Using a panel of deletion mutants, we found that TGF-beta binds to the NH2-terminal endoglin-related region of betaglycan. The remainder of the extracellular domain and the cytoplasmic domain are not required for presentation of TGF-beta to receptor II; however, membrane anchorage is required. Soluble betaglycan can bind TGF-beta but does not enhance binding to membrane receptors. In fact, recombinant soluble betaglycan acts as potent inhibitor of TGF-beta binding to membrane receptors and blocks TGF-beta action, this effect being particularly pronounced with the TGF-beta 2 isoform. The results suggest that release of betaglycan into the medium converts this enhancer of TGF-beta action into a TGF-beta antagonist.