You are on page 1of 15

Review

Extracellular matrix contribution to


skin wound re-epithelialization

Patricia Rousselle, Marine Montmasson and Cécile Garnier


Laboratoire de Biologie Tissulaire et Ingénierie Thérapeutique, UMR 5305, CNRS - Université Lyon 1, Institut de Biologie et Chimie des
Protéines, SFR BioSciences Gerland-Lyon Sud, 7 passage du Vercors, F-69367, France

Correspondence to Patricia Rousselle: patricia.rousselle@ibcp.fr.


http://dx.doi.org/10.1016/j.matbio.2018.01.002

Abstract
The ability of skin to act as a barrier is primarily determined by cells that maintain the continuity and integrity of skin
and restore it after injury. Cutaneous wound healing in adult mammals is a complex multi-step process that
involves overlapping stages of blood clot formation, inflammation, re-epithelialization, granulation tissue formation,
neovascularization, and remodeling. Under favorable conditions, epidermal regeneration begins within hours after
injury and takes several days until the epithelial surface is intact due to reorganization of the basement membrane.
Regeneration relies on numerous signaling cues and on multiple cellular processes that take place both within
the epidermis and in other participating tissues. A variety of modulators are involved, including growth factors,
cytokines, matrix metalloproteinases, cellular receptors, and extracellular matrix components. Here we focus on
the involvement of the extracellular matrix proteins that impact epidermal regeneration during wound healing.
© 2018 Elsevier B.V. All rights reserved.

Introduction macrophages, which enter the wound slightly later [4].


Macrophages also secrete growth factors that attract
The ability of skin to act as a barrier is primarily keratinocytes, fibroblasts, and blood vessels into the
determined by cells that maintain the continuity and wound, thereby promoting re-epithelialization, granu-
integrity of skin and restore it after full-thickness lation tissue formation, and vascularization (Fig. 1B)
wounds (Fig. 1A). Cutaneous wound healing is not, as [3,5]. Both keratinocytes and fibroblasts synthesize
was long believed, a simple linear process in which key ECM components that are needed to re-form the
growth factors are synthesized that activate cell basement membrane (BM) beneath the epidermal
proliferation and migration. Rather, wound healing is basal layer. There is wound contraction due to
a complex multi-step process that involves closely fibroblasts pulling on the ECM; this facilitates human
related dermal and epidermal events. Skin repair is wound remodeling, and the associated traction forces
the result of dynamic and interactive processes that contribute to scarring [3]. The orchestration of these
involve soluble factors, blood elements, extracellular events relies on the spatial and temporal expression
matrix (ECM) components, and cells [1]. In adult and the activation of a variety of proteins, such as
mammals, skin repair can be divided into four growth factors, cytokines, matrix metalloproteinases
continuous phases, namely hemostasis, inflammation, (MMPs), and ECM components. In contrast, wounds in
proliferation (including re-epithelialization, granulation mammalian embryos heal via rapid re-epithelialization
tissue formation, and neovascularization), and matu- and, in the absence of inflammation, via granulation
ration or remodeling, and it generally leaves a scar tissue formation and contraction [6].
[2,3]. In mammals, blood clotting occurs first, and this
provides chemotactic factors that attract inflammatory Epithelialization of wounds
leukocytes and ECM proteins that serve as a substrate
for migrating cells. Neutrophils help cleanse the Re-epithelialization is a critical step in the
wounded area and are eventually phagocytosed by wound healing response (Fig. 1B). Defects in

0022-2836/© 2018 Elsevier B.V. All rights reserved. Matrix Biol. (2019) 75-76, 12–26
Extracellular matrix contribution to skin wound re-epithelialization 13

Fig. 1. Diagram showing the re-epithelialization process. (A) Schematic drawing of the epidermis resting on the basement
membrane (BM). Mammalian skin consists of the epidermis and dermis, which are separated by a BM. The epidermis is a
stratified squamous epithelium that is composed of several cell layers, with the basal layer resting on the BM. The diagram
shows adhesion structures, anchoring complexes, and BM organization. a, laminin 332; b, laminins 511/521; c, perlecan;
d, collagen XVII; e, collagen VII; f, nidogen; g, collagen IV; h, syndecan; i, collagen I/III. (B) The re-epithelialization process 1 to
5 days post-wounding. Approximately 24 h after injury, activated basal and suprabasal keratinocytes adjacent to the wound
edges undergo dramatic morphological changes. They acquire a motile phenotype and start migrating over the granulation
tissue to cover the wounded area. Induction of a new set of genes leads to the expression of cellular receptors, growth factors,
cytokines, matrix metalloproteinases and extracellular matrix proteins. Keratinocytes behind the wound edges begin
proliferating and enrich the migrating tongue. The provisional matrix and granulation tissue are indicated.

re-epithelialization lead to chronic non-healing clot [5]. In this review, we focus on ECM proteins
wounds, such as venous stasis, diabetic and pressure involved in the re-epithelialization of wounds. Notably,
ulcers [7]. Under favorable conditions, epidermal the involvement of cytokines, growth factors, and
regeneration begins within hours after injury and MMPs has been described elsewhere [13,17,18].
continues until the epithelial surface is intact.
The inflammatory reaction to wounding results in the
appearance of matrix molecules in the wound bed, The provisional ECM during epidermal
with keratinocytes migrating rapidly from the edges of repair
the wound and from hair follicles and sweat glands into
the wound bed (Fig. 1B) [8,9,10]. As fibroblasts begin The migrating keratinocytes are in close contact
to produce new ECM, epithelialization is initiated by with a provisional ECM that is associated with early
epidermal growth factor (EGF) and transforming wounds and is mainly composed of fibrin, plasma
growth factor (TGF) α, which is produced by platelets, fibronectin (FN), vitronectin, and platelets [15,19,20].
macrophages, and keratinocytes [11,12,13]. There are two forms of FN: plasma FN, which is
Normally, the epithelialization of wounds involves synthesized in a soluble form by hepatocytes into the
an orderly series of events in which keratinocytes blood plasma [21], and cellular or tissue FN, which
migrate, proliferate, and differentiate to restore the is produced by other cells, such as fibroblasts,
barrier function [14]. When the epithelium has covered endothelial cells, and keratinocytes [22] (Fig. 2A).
the wound bed, the proteins of the dermal-epidermal Plasma FN is more important in the early phase of
junction (DEJ) reappear in sequential order from the wound healing when it binds to platelets and fibrin,
margins to the center of the wound, due to the con- thereby increasing the strength of the fibrin clot
certed action of epidermal and dermal cells (Fig. 1A) [21,23,24]. The strong adhesion of keratinocytes to
[15,16]. Basal keratinocytes then revert to a stationary the BM is disrupted when their hemidesmosomes
phenotype in which they have apical polarity and dissolve; when their cell-cell connections are weak-
are firmly anchored to the DEJ through reconstituted ened, they then start to migrate underneath or through
hemidesmosomes on their basal surface. the temporary scaffolding that fills the wound defect
Epithelialization steps are stimulated by the local [25]. In vivo studies have shown that plasma FN plays
wound milieu, which typically shows altered ECM an important role in keratinocyte adhesion and
composition and the presence of cytokines that are migration [26]. Migrating keratinocytes are highly
produced by cells in the granulation tissue and by the phagocytic, which allows them to penetrate through
14 Extracellular matrix contribution to skin wound re-epithelialization

A C 5 LN
Type I Type II Type III
Plasma Fibronectin :
V
NH2 COOH
SS LEa

Cellular Fibronectin : EDA V domains


NH2 COOH
SS LF
EDB V LEb
NH2 COOH
SS L4
B 3A
? Cleaving LEc
LE 1/ 2
3 2 enzymes: 1
mTLD
S S S S S S S S S
LN LE S S S LEb L4 LEa S S S
BMP-1 LN LEa LF LEb S S S LEb L4 LEa LN
domains domains domains
domains

Coiled-coil
Coiled-coil

L
L

Coiled-coil
S S
H Maturation H

S
H

Cleaving
enzymes:

S S Binding sites for : S S plasmin S S


1 mTLD
1 Integrin 3 1, 6 1, 6 4 1
LG 2 3 2 3 BMP-1 LG 2 3
4 5 Syndecan-1 and -4 thrombin 4 5
MMP?

Fig. 2. Schematic representation showing human fibronectin (FN), laminin (LM) 332, LM 511 and LM 521. (A) FN is a multi-
domain glycoprotein consisting of a series of 3 different modules, namely FNI, FNII, and FNIII repeats. These serve as building
blocks for structural domains that have different functions [183]. For example, FNI are found in fibrin binding domains, FNII are
found in collagen binding domains, and FNIII are found in cell-binding domains [21]. Three alternatively spliced regions include
the extra domain (ED) A and EDB regions, both type III repeats, and a variable (V) region [184]. Alternative splicing of the FN
gene is important in the context of wound repair, since alternatively spliced forms are present in the matrix due to migrating
keratinocytes as well as in the granulation tissue [24,42,44]. Insertion of additional type III repeats, EIIIA and EIIIB, which flank
the major cell adhesion domain of the molecule results in FNs EDA and EDB, respectively. EDA and EDB are present in
so-called cellular FN but are absent from plasma FN. FN is secreted as a dimer in which the monomers are linked by disulfide
bonds, which is essential for FN matrix assembly. (B) The structure of human LM 332 and its physiological maturation process.
LM 332 is composed of three subunits: α3A, β3, and γ2. Each chain has the indicated domains. The large LG structure located
at the C-terminal end of the α3 chain contains five repeating LG domains. The first three repeats, LG1–3, interact with the α3β1,
α6β1, and α6β4 integrins, while LG4 and LG5 contain binding sites for syndecan-1 and -4. LM 332 is synthesized as a
precursor molecule that undergoes maturation by proteolytic processing at the α3Α chain N- and C-terminus as well as at the
γ2 chain N-terminus. The enzymes that are known to be involved are shown, and the cleavage sites are indicated by arrows.
(C) Structures of human LM 511 and LM 521.

the clot. Degradation of the fibrin-FN clot appears to the fibrin matrix remain unknown, and it is also
be critical for wound healing, since mice that lack the unknown whether this migration process takes place
plasminogen gene or the urokinase-type plasminogen in vivo, where keratinocytes migrate collectively. The
activator/tissue-type plasminogen activator gene migration of keratinocytes through a clot may be the
exhibit a number of wound abnormalities, including result of the synergistic action of MMPs and plasmin.
defects in wound healing and in keratinocyte migration FN plays a role in inflammation by promoting macro-
[27,28,29]. phage activation and opsonizing debris [15]. Kerati-
Interestingly, keratinocytes migrate through tunnels nocytes deposit newly produced ECM proteins along
of digested fibrin matrix in vitro, a process that requires their migration path, including cellular FN, tenascin C,
plasminogen activation at the leading edge [30]. and laminin (LM) 332 (Fig. 3) [8,31,32,33], suggesting
Tunnel formation and cell track length are differentially that wound edge keratinocytes can modulate their own
influenced by cytokines such as TGFβ1 and EGF. The adhesion and migration substrates. Surprisingly, the
mechanisms underlying keratinocyte adherence to wound healing process, including re-epithelialization,
Extracellular matrix contribution to skin wound re-epithelialization 15

Epidermis Soluble matrix molecules


and fragments: Provisional
Laminin 3 LG45 Matrix:
Tenascin C Plasma Fibronectin
Fibronectin Fibrin
Vitronectin
Tenascin C

Reforming basement
Membrane:
Laminin 332
Laminins 511/521 Epithelial tongue:
Collagen IV Precursor laminin 332
Perlecan Fibronectin EDA
Nidogens Tenascin-C
Collagen VII Collagen VII
Collagen XVII Early granulation tissue:
Collagen Collagen VI
Collagen V - Proteoglycans
Fibronectin EDA and EDB
Thrombospondin1
Osteopontin, SPARC
Tenascin C

Fig. 3. Schematic showing the localization of extracellular matrix (ECM) components during re-epithelialization. The major
ECM proteins are indicated in the vicinity of the epithelial tongue during epidermal regeneration 1 to 3 days post-injury (other
extracellular molecules i.e. matrix metalloproteinases, cytokines, growth factors an ECM receptors are not represented here).

appears normal in plasma FN-deficient animals [34], type-specific manner, and the splicing variant can
suggesting that cellular FN may compensate for the determine the functional properties of the molecule. In
lack of plasma FN. As the provisional matrix matures, humans, there are over 20 main isoforms of FN that
cell-mediated remodeling takes place that is directed are generated after mRNA processing, while rats and
primarily by inflammatory cells and fibroblasts. This mice produce 12 variants [40,41].
remodeling includes a transition from a fibrin-FN- Alternative splicing of the FN gene transcript is
dominated scaffold to a tissue-FN-dominated one that important in the context of wound repair [24,42].
contains proteoglycans and interspersed crosslinked In situ hybridization studies of skin wound healing in
collagen; the tissue progressively becomes granula- adult rats show that wounding induces FN mRNA
tion tissue with which keratinocytes interact (Fig. 1B) expression directly, both beneath the wounded area
[23,35]. and in adjacent dermal tissue. This increased level of
FN mRNA is evident within 1 day after wounding,
with the highest levels seen on days 4 and 6 after
Cellular fibronectin wounding. The splicing pattern of FN mRNA in the
wound bed differs from that in healthy skin. Cells at
FN is an important constituent of the vertebrate the base and edges of the wound express mRNA
ECM. It is ubiquitously expressed in human tissue and that includes both the EDA and EDB regions; in
it has long been known to be a major component in contrast, mRNA in the adjacent dermis and muscle
wound healing [36,37]. FN is a multi-domain glyco- shows the same splicing pattern as in normal skin
protein composed of an array of modular structures and lacks the EDA and EDB regions [24]. After
(Fig. 2A). It is secreted as a dimer in which the wounding, macrophages are the first cells to show
monomers are linked by a pair of disulfide bonds at the increased FN mRNA expression, which can be seen
C-terminus [21,38]. The FN gene has three alterna- as early as 2 days after the injury. Notably, some of the
tively spliced domains: extra domain (ED) A or extra macrophage FN mRNAs contain EDA and EDB
type III homology (EIII) A, EDB or EIIIB, and the type domains [42,43]. Later, 7 days after wounding, gran-
III homologies connecting segment (IIICS; V region ulation tissue fibroblasts begin to express FN proteins,
in rat) [39]. FN alternative splicing varies in a cell both with and without the EDA and EDB domains
16 Extracellular matrix contribution to skin wound re-epithelialization

[42]. Additional studies revealed that in humans, when process, showing collagen I mRNA in the upper layer
keratinocytes migrate into wounds, they deposit of granulation tissue 6–13 days after wounding [59]. In
EDA-containing FN underneath the leading epithelial early granulation tissue, there are high levels (30%) of
tongue, while both EDA- and EDB-containing FN collagen III, whereas its expression level in the mature
isoforms are found in the granulation tissue (Fig. 3) scar is the same as in the intact dermis (10%) [60].
[44,45]. In vitro, migrating cultured keratinocytes leave Even though the migration tongue progresses signif-
behind trails of EDA-containing FN that correlates with icantly before the appearance of collagen I in the
their migration [44]. Transgenic mice that lack the EDA granulation tissue, a number of studies suggest a
exon show major defects in re-epithelialization during potential role for collagen I in re-epithelialization. For
the late phase of wound healing (days 5–7 after example, one study reported that keratinocytes need
wounding), consistent with the observation that in the to be activated by collagen I integrins before they
first days after wounding, deposited FN is derived migrate [61], and another study revealed that kerati-
mainly from plasma [46]. In addition to epidermal nocytes that are in direct contact with collagen I begin
ulceration, the granulation tissue in EDA-deficient to increase their production of collagenase [62].
mice shows edematous regions that lead to abnormal Ligation of integrin α2β1 to dermal matrix collagen I
wound healing. Although no wound healing defects stimulates the expression and catalytic activity of
are seen in mice that lack the EDB segment, in vitro MMP-1 in keratinocytes, thus facilitating directional
studies reveal reduced fibroblast proliferation and keratinocyte migration over the dermal matrix [63].
defects in FN matrix assembly [47]. Taken together, While MMP-1 is present in human cutaneous wounds
these data suggest that the EDA and EDB domains during re-epithelialization, it is no longer expressed
have distinct functions. when wound closure is complete [64]. A study that
The N-terminal portion of FN contains a self- showed the capacity of collagen I to promote
assembly domain that allows FN dimers to assemble keratinocyte migration in vitro suggested that it was a
in a process termed fibrillogenesis, which produces major substrate for keratinocyte migration during
a three-dimensional (3D) matrix [39]. In contrast to epithelialization [65]. Additional studies suggested
collagen self-assembly, FN fibrillogenesis is a non- that keratinocytes do not invade the fibrin-rich clot in
spontaneous, cell-mediated process [48]. FN is vivo; rather, they migrate on a pathway between the
secreted in a compact conformation, and it must fibrin clot and the collagen-rich viable dermis using
interact with cell-surface integrins, most notably dermal collagen I as a substrate for their migration
the α5β1 integrin, in order to reveal the cryptic self- (Fig. 3) [15,66]. Further work is needed to elucidate
association domains needed to form a definitive 3D whether and how collagen I impacts epidermal
matrix [49]. FN polymerization is a critical regulator closure. A recent study unexpectedly challenged the
of ECM organization and stability. Specifically, it is notion that fibrillar collagens are recognized by
essential for the deposition and retention of collagen I collagen-binding integrins [67].
and thrombospondin (TSP) 1 within the ECM, and it Collagen V has been implicated in the dermal wound
regulates the turnover and endocytosis of collagen I, healing response in terms of modulating fibroblast
which stabilizes collagen I matrix fibrils in granulation behavior and tissue contraction [68]. Patients with
tissue [50,51]. Interestingly, a sub-population of classical Ehlers-Danlos syndrome, which is caused
macrophages appears to participate in collagen by collagen V gene mutations, show skin fragility and
endocytosis and degradation [52]. abnormal wound healing [69]. Additional studies in
mice show that reduced collagen V expression is
associated with decreased fibroblast proliferation and
Collagens migration and with reduced attachment to the ECM
[70,71].
Dermal reconstitution begins approximately 3 to Non-fibrillar collagen VI is also found in granulation
5 days after injury and is characterized by granulation tissue. It is deposited during the early stages of wound
tissue formation, which involves new blood vessel healing, and its synthesis increases in parallel with
formation and the accumulation of fibroblasts. As collagen I, reaching a peak about 8 days post-
dermal fibroblasts begin to migrate into the wound wounding [72,73]. A recent study has shown that
site, fibrillar collagen production gradually increases. collagen VI is a key regulator of dermal matrix
The resulting collagen matrix ultimately replaces the assembly and composition and that it influences
provisional fibrin matrix [53]. Collagen comprises fibroblasts behavior [74]. Interestingly, macrophages
more than 50% of the protein in scar tissue, and its were shown to produce collagen VI to modulate their
production is essential to the healing process [54]. own binding properties [75].
Collagen III and FN are deposited in the initial steps of Recent studies have revealed that other collagens
healing; later on, collagen III is gradually replaced by may play roles in the repair process by directly or
collagen I, the deposition of which can be observed on indirectly influencing the re-epithelialization process.
day 7 after wounding [55,56,57,58]. In situ hybridiza- An early analysis of collagen VII expression did not
tion experiments have confirmed this replacement support its potential involvement in skin
Extracellular matrix contribution to skin wound re-epithelialization 17

epithelialization [8]; however, more recent studies of hypertrophic scar model reduced the size of scars,
genetic recessive dystrophic epidermolysis bullosa possibly by reducing vascularization and angiogen-
(RDEB) disorders revealed that the loss of collagen esis [85].
VII results in delayed re-epithelialization by altering
keratinocyte migration that is associated with the
irregular deposition of LM 332. Moreover, the Laminins
absence of collagen VII in skin delays fibroblast
migration and granulation tissue formation [76]. These All LMs are composed of three different gene
findings reinforce previous studies that showed that products, termed the α, β, and γ chains, that are
collagen VII plays a role in keratinocyte migration [77]. assembled into an αβγ heterotrimer. LM α chains
Collagen VII is found in the provisional ECM under the have a large LM globule, termed LG, at the
epidermal tongue [76]. After topical application of carboxy-terminal end that consists of 5 structurally
recombinant collagen VII to full-thickness wounds in similar domains (LG1 to LG5) that each contains
a murine wound healing model, it became stably about 200 residues (Fig. 2B) [86]. The major LM
incorporated into the DEJ and it accelerated wound isoforms that are expressed in the DEJ, namely LM
closure by increasing re-epithelialization and decreas- 332 (formerly called kalinin, epiligrin, and nicein)
ing scarring and pro-fibrotic markers [78]. Applying [87], LM 511 and LM 521 appear to be involved in
collagen VII to collagen VII-null skin that was grafted wound repair (Fig. 2B, C). Increased LM 332
onto athymic nude mice corrected the poor dermal expression is one of the earliest events in wound
epidermal adherence and the anchoring fibril defects epithelialization, and LM 332 is expressed in
[78]. epidermal keratinocytes within hours of injury. LM
In vivo, collagen IV is in turn, not highly deposited in 332 appears to be the first BM component laid down
the provisional matrix under migrating keratinocytes. onto the wound bed, as its expression precedes that
Immunohistochemical analysis of human skin wounds of all other ECM components (Fig. 3)
reveals its expression in BM fragments as early as [88,89,90,91,92]. The molecular profiling of leading
4 days after wounding. After 8 to 21 days, collagen IV edge keratinocytes in a mice wound model identified
can be detected in the newly formed BM [8,79], LM 332 and FN as the major ECM proteins expressed
confirming its role in BM assembly and stabilization. by these cells [93]. LM 332 is synthesized by
Transmembrane collagen XVII, also known as the keratinocytes as a 460-kDa high molecular weight
hemidesmosomal component bullous pemphigoid precursor protein. After secretion and deposition into
antigen (BPA) 180, plays a role in keratinocyte the ECM, the α3 and γ2 chains undergo maturation
adhesion to collagen IV and migration in vitro [80]. events consisting of loss of the 2 globular LG45
Further, its increased expression and the shedding domains at the C-terminus of the α3 chain and loss of
of its collagenous ectodomain in acute skin wounds a 50-kDa fragment at the N-terminus of the γ2 chain
suggest that it might play a role in keratinocyte motility (Fig. 2B) [94]. These physiological post-translational
and proliferation [81]. Analysis of wound closure in mice processing events appear to allow the stable integra-
that expressed a functional non-shedding collagen tion of LM 332 into the BM [95,96]. Mature LM 332 is
XVII revealed that collagen XVII ectodomain shedding the major component of anchoring filaments in skin
is a highly dynamic modulator of in vivo proliferation [97] where it mediates cell adhesion via interaction of
and motility in activated keratinocytes during epidermal the α3 carboxyl-terminal LG1–3 triplet domain with
regeneration [81,82]. Further analysis revealed a role both α3β1 and α6β4 integrins [98,99,100,101], while
for collagen XVII in keratinocyte adhesion and direc- the N-terminal short arms connect to BM components
tional motility through integrin-dependent PI3K activa- (Fig. 2B). In vivo, epidermal injury activates the
tion and through stabilization of the lamellipodia at transcription and deposition of LM 332 into the
the leading edge of re-epithelializing wounds [83]. provisional matrix by the leading keratinocytes in the
The heparan sulfate proteoglycan (HSPG) colla- process of epidermal outgrowth and migration at the
gen XVIII is also suspected to play a role in wound wound edge [88,102]. Precursor LM 332 is found in
healing. Genetically modified mice that lack collagen this provisional matrix, but it is absent from mature
XVIII show accelerated wound healing [84]. In these BMs [103,104]. Keratinocytes leave behind trails of
mice, the vascularization rate is accelerated in the precursor LM 332 while migrating on collagen I [91] or
wound bed, and the myofibroblast density is after stimulation by TGFß1 [105]. Because LM 332
increased, suggesting that the accelerated healing with a full-length α3 chain is found in epidermal repair
is due to changes in wound contraction. Conversely, [89,106,107], it was assumed that the LG45 domains
wound healing is delayed in mice that overexpress also play a role (Fig. 2B). Indeed, a function for LG45
the collagen XVIII carboxy-terminal endostatin do- in the deposition of LM 332 in the ECM was
main in their skin; these mice have abnormalities in hypothesized [94,108], as was a direct function in
the epidermal and endothelial BM, but show no keratinocyte adhesion and migration [107,109,110].
significant effects on angiogenesis [84]. On the other Syndecan-1 and -4 have been identified as cellular
hand, the use of endostatin in a rabbit ear receptors for the LG45 domains [111,112], supporting
18 Extracellular matrix contribution to skin wound re-epithelialization

the idea that keratinocytes migrate by forming Perlecan


actin-based cellular protrusions [113] or by activating
MMP-1 (Fig. 2B) [111]. It has been shown that the Perlecan has a structural bridging function in the
expression and processing of the LG45 module is BM. It helps assemble the major constituents of the
enhanced in keratinocytes after infection and in BM, including collagen IV, LMs, and nidogen/
chronic wounds and that the level of expression and entactin [128,129], within molecular suprastruc-
further processing correlate with the speed of wound tures, thereby creating molecular bridges that fit
healing [114]. Furthermore, LG45-derived peptides with cellular interactions [130]. Perlecan is the
have antimicrobial activity and chemotactic activity in major proteoglycan component of BMs [131,132].
mononuclear cells, suggesting that LG45 may also As such, it acts as a reservoir for heparin-binding
participate in host defense [114]. The full-length γ2 growth factors, limiting their diffusion and therefore
chain is also an important regulator of keratinocyte their ability to act on cells on either side of the BM
migration during wound repair [105,115,116], and it is [133]. The degradation of perlecan during wound
hypothesized to be involved in LM 332 incorporation healing or tissue remodeling may allow the rapid
during BM assembly [117]. The sequential analysis of introduction of mitogens and trophic factors to
partial-thickness wounds on the forearm skin of begin the regenerative processes. It can also
healthy human individuals revealed that the γ2 is impact cells through direct interactions [133]. One
expressed continuously by migrating keratinocytes study reported that the BM HSPG perlecan
during reepithelialization while collagens IV and VII deposition in the ECM of confluent keratinocytes
are deposited after wound closure [118]. strengthens cell adhesion and inhibits keratinocyte
The α3β1 integrin was identified as an important motility [123]. Furthermore, perlecan co-localizes
mediator of LM γ2 processing, both in vivo during with LM 332 at the wound margin of full thickness
wound healing and in vitro in cultured keratinocytes wounds in mice and is concentrated in areas in
[119] where it acts by upregulating the mammalian which the BM matures to stabilize keratinocyte
tolloid (mTLD)/bone morphogenetic protein-1 (BMP-1) adhesion (Fig. 3). This finding suggests the
metalloproteases (Fig. 2B) [120]. In contrast, the proximity of these two components during BM
α9β1 integrin plays a role in the inhibition of LM γ2 remodeling [123,130]. Studies in perlecan heparan
processing by antagonizing α3β1-dependent mTLD/ sulfate-deficient mice revealed that the mice had
BMP-1 expression during re-epithelialization [120]. significantly delayed wound healing due to reduced
Studies of wound healing after conditional knockdown granulation tissue formation as well as decreased
of the activity of BMP-1/tolloid-like proteinases in mice vascular density, although epidermal closure was not
reveal delays in re-epithelialization and a dramatic affected [134]. In line with its potential involvement in
reduction in ECM deposition [121]. The underlying the late phases of epidermal regeneration, perlecan
mechanism most likely involves incomplete process- plays a role in keratinocyte survival and terminal
ing of precursor ECM proteins, including that of LM γ2 differentiation [135]. More recently, it was shown to
[121]. influence keratin 15 expression in keratinocytes [136],
LM 511 and LM 521 may also contribute to suggesting that it may influence keratinocytes stem-
epidermal regeneration (Fig. 2C). Although they ness during wound repair.
are described as potent keratinocyte adhesion
substrates with the capacity to induce proliferation
and migration in vitro [122], they are not found under Nidogens
migrating keratinocytes [123,124], and their expres-
sion and deposition in the maturing BM increases The BM isoforms nidogens 1 and 2 bind to several
after re-epithelialization is complete [76,123]. Their BM-associated proteins. They have been described
expression correlates with the increased ability of as connecting elements between the LM and
keratinocytes to regenerate intact epithelium, and collagen IV networks that facilitate BM formation
detachment from LM 511/521 in the BM is likely to be and stabilization [137,138,139]. Mice that lack both
a key factor in promoting keratinocyte differentiation isoforms die shortly after birth from numerous tissue
[122]. LM 511 is an abundant isoform, both at abnormalities related to defects in BM assembly
interfollicular sites and in the surrounding hair [140], but analysis shows that their skin has an
follicles, and it is synthesized by basal keratinocytes ultrastructurally normal BM [141]. This demonstrates
[125]. In a regenerative context, LM 511 is described that some BMs can form in vivo without nidogen.
as also being produced by dermal pericytes that are Nidogen 1 was proposed to play a non-structural role
adjacent to the proliferative basal layer of the in skin wound repair [142]. Wound healing studies in
epidermis [126]. Recent data from a nidogen 1-deficient mice revealed no defects in
keratinocyte-specific LM α5 KO suggest that LM α5 re-epithelialization, but the newly formed epidermis
inhibits keratinocyte proliferation and migration by showed marked hyperproliferation and delayed
controlling epidermal and dermal growth factor differentiation with an altered deposition of LM γ1
expression [127]. chain.
Extracellular matrix contribution to skin wound re-epithelialization 19

Matricellular proteins part of an intracellular signaling complex, depending on


the cellular context. A number of studies have reported
Matricellular proteins are multi-domain extracellu- elevated OPN expression in fibrosis, supporting the
lar proteins. They do not appear to play structural idea that OPN may contribute to inflammation-
roles in tissues. Rather, they are involved in a wide associated fibrosis. Even though OPN is needed for
variety of cellular processes and are also able to proper wound healing, it leads to fibrosis and excess
influence cell behavior associated with wound scar formation if it persists for too long or in too high a
healing. Many matricellular proteins show increased concentration. OPN acts as a cytokine that is a
expression in response to injury, including TSP1 and chemoattractant for fibroblasts, and it appears to be
TSP2, osteopontin (OPN), tenascin C, and SPARC necessary for fibroblast deposition of ECM compo-
(secreted protein acidic and rich in cysteine) [143]. nents and collagen (Fig. 3) [153]. There are changes in
the matrix architecture and in collagen fibril formation in
the deep layers of wounds in OPN-deficient mice,
Thrombospondins indicating that OPN is involved in matrix reorganization
[154]. One study reported faster re-epithelialization and
TSP1 and TSP2 are matricellular proteins; as such, decreased granulation tissue/scar formation in the
they are released and/or secreted following tissue healing of skin wounds of mice with acute local
injury and exert their cellular effects by associating knockdown of OPN [155]. Macrophage-derived platelet
with the ECM. While TSP1 and 2 are produced by derived growth factor-BB appears to be responsible for
fibroblasts, TSP1 is a major component of platelet inflammation-induced OPN expression in wound fibro-
α-granules and is found in the circulation. TSP1 and 2 blasts, indicating that inflammation-triggered expres-
appear to play distinct, non-overlapping roles in the sion of OPN both hinders the rate of re-epithelialization
healing of skin wounds. Although TSP1 and 2 are and contributes to wound fibrosis.
dispensable for developmental angiogenesis, they
are critical for proper tissue repair, both directly
through their interactions with cell receptors and/or Tenascin-C
indirectly by modulation of the levels or bioavailability
of effector molecules, including growth factors and Tenascin-C is a large hexameric extracellular
MMPs [144]. In addition, TSP1 is involved in the glycoprotein. The founding member of a family of
activation of latent TGFβ1 [145] and is considered a four tenascins, it has a unique and distinct expression
major in vivo activator of TGFβ1, as TSP1-null mice pattern. Tenascin-C binds many extracellular partners,
have lower active TGFβ1 levels [146]. The spatial and including matrix components and soluble factors; it
temporal expression of TSP1 and 2 are distinct in also influences cell phenotype through interactions
dermal wound healing [147,148]. TSP1 is induced with cell surface receptors.
early in wound healing, concomitant with thrombus Little or no tenascin-C is detected in healthy adult
formation, edema, and inflammation. TSP2 expres- tissues. After tissue damage, tenascin-C plays a
sion occurs after resolution of the inflammatory phase multitude of roles that mediate both inflammatory
and during granulation tissue remodeling. Wound and fibrotic processes to enable effective tissue repair
healing studies in TSP1-null or knockdown mice [156]. Tenascin-C mRNA and protein expression is
indicate delayed repair that is characterized by loosely induced rapidly upon skin injury [72,157,158]. De novo
compacted and disorganized granulation tissue due tenascin-C expression is a hallmark of inflammation,
to decreased levels of both total and active TGFβ1 and a number of cytokines can induce the gene
and delayed re-epithelialization [148,149]. Full thick- [159,160]. It is expressed in areas with increased
ness wounds in TSP2-null animals heal at an immune cell infiltration during the acute inflammation
accelerated rate compared to WT mice and show phase, and it co-localizes with polymorphonuclear
alterations in the remodeling phase of healing [150]. lymphocytes in the inflamed human dermis [161].
Complete re-epithelialization of full thickness exci- Tenascin-C is suggested to have a role in stimulating
sional wounds occurs earlier than in WT mice, and the lymphocyte migration and activation [162,163]. It
epithelial layer is thicker. Further, granulation tissue at stimulates the production of the pro-inflammatory
the wound site is disorganized: the morphology of cytokines tumor necrosis factor α, interleukin (IL)-6,
collagen fibers is altered as is FN distribution; these and IL-8 in primary human macrophages and synovial
effects are most likely secondary to increased MMP-2 fibroblasts via activation of Toll-like receptor
and MMP-9 expression [151]. 4-mediated signaling pathways [164]. In addition its
expression is also observed within 24 h of injury
underneath migrating and proliferating keratinocytes
Osteopontin [72,165]. In situ hybridization demonstrates that
keratinocytes are the major source of tenascin-C
OPN is a phosphorylated acidic glycoprotein [152] (Fig. 3) [166], suggesting that tenascin-C may help
that acts as a secreted chemokine-like protein and as promote the proliferation and migration of epithelial
20 Extracellular matrix contribution to skin wound re-epithelialization

cells. Tenascin-C is increased in the wound edges of phages start the process by providing growth
the dermis associated with activated fibroblasts and factors, chemokines, cytokines and ECM proteins
extends throughout the dermis. High levels are also that activate the proliferation phase; this phase is
found in the new granulation tissue [72,157,165,167], beneficial and necessary for the timely repair of
suggesting a role for tenascin-C in and promoting injured skin [175]. Besides, studies have revealed
fibroblasts migration and tissue rebuilding during that macrophages exert specific functions during the
wound healing. Tenascin-C expression is not detected different stages of skin repair [176]. The conditional
after wound repair is complete; indeed, tenascin-C depletion of macrophages in specific phases of the
levels return to normal after wound contraction, and repair response in mice revealed that early-stage
the protein is not present in scar tissue [72,165,167]. macrophages control induction of granulation tissue
The persistent expression of tenascin-C is concomi- and myofibroblast differentiation, while mid-stage
tant with chronic inflammation, and it is persistently macrophages play a role in stabilization of vascular
upregulated in fibrotic disease, including keloids and structures and transition of granulation tissue into
photodamaged skin [168,169]. scar tissue [176]. Macrophages are also a source of
ECM proteins, and they express high levels of
proteases that contribute to ECM cleavage and
SPARC remodeling [177,178]. Concomitantly, these early
inflammatory signals appear to influence fibrosis
SPARC, also known as BM-40 and osteonectin, is a and scar formation [18]. Interestingly, studies have
32-kDa calcium-binding glycoprotein that is secreted revealed that wounds in the epidermis of zebrafish
by various cells, including fibroblasts, endothelial cells, and mammalian embryos close via similar princi-
and platelets [170]. Although SPARC is expressed in ples [6]. But, in these systems, wounds are
both the dermis and epidermis of uninjured skin, its re-epithelialized extremely rapidly and independently
expression increases during wound healing early after of inflammation and ECM proteins supplied by the
injury and reaches the highest level later during blood clot or granulation tissue [179]. This might be
remodeling of the dermal ECM [171,172]. Interestingly, due to a predominant impact of tissue-autonomous
SPARC-null mice display accelerated cutaneous extension movements within the re-epithelializing
wound closure [173], which could be due to the epidermis, a process that is likely to also occur in
reduced collagen content of the SPARC-null dermis wounds of mammalian embryos [180]. In adult zebra
that appears to render the skin intrinsically more fish, keratinocytes repopulate the wounds via TGFβ-
susceptible to cell contraction. In addition to modulating dependent and integrin-dependent lamellipodial
the activity of a number of growth factors and ECM, crawling at the leading edges of the epidermal
SPARC may also influence MMP expression by tongue, thereby leading to epidermal closure that
several types of cells that participate in wound healing. is independent of keratinocyte proliferation [181].
This process resembles the primary healing or
resealing process in small or acute wounds in adult
Proteoglycans and glycosaminoglycans mammals, where the wound edges are very close.
(GAG) The actions of myofibroblasts are minimized and
the regeneration of the epidermis is optimized, since
Proteoglycans and GAGs play key roles in both epidermal cells only need to migrate a very short
fibrotic and regenerative wound healing. Their involve- distance [182].
ment has been described mainly in dermal repair Future research will undoubtedly clarify which
processes, scar formation, and ECM remodeling [174]. inflammatory cues and which ECM proteins (either
produced by the keratinocytes themselves or present
within the granulation tissue) are central to efficient
Final considerations re-epithelialization in adult mammals. This will identify
putative targets for innovative therapeutic strategies
Several ECM proteins play roles in wound re- that favor wound closure in clinical situations in which
epithelialization, either via direct contact with kera- healing is delayed or prevented.
tinocytes or indirectly by affecting the granulation
tissue. The quality of the granulation tissue, as well
as its sequentially modified composition and struc-
ture, clearly influence the speed of epithelialization
and the properties of the newly produced epidermis. Acknowledgements
While disorganized or excessive granulation tissue
correlates with delayed re-epithelialization, de- Work in the authors' laboratory was supported by
creased granulation tissue formation and/or en- the Agence Nationale de la Recherche (grant number
hanced contraction speeds re-epithelialization. In ANR-13-RPIB-0003-01), by the Ligue Nationale
adult mammals, inflammatory cells such as macro- Contre le Cancer (C119235) and by CNRS in the
Extracellular matrix contribution to skin wound re-epithelialization 21

framework of the GDR GAG (GDR 3739). MM and CG cutaneous wounds, Wound Repair Regen. 21 (2013)
were both funded by the Agence Nationale de la 194–210, https://doi.org/10.1111/wrr.12029.
Recherche (grant number ANR-13-RPIB-0003-01). [8] H. Larjava, T. Salo, K. Haapasalmi, R.H. Kramer, J. Heino,
Expression of integrins and basement membrane compo-
nents by wound keratinocytes, J. Clin. Invest. 92 (1993)
Received 30 September 2017;
1425–1435, https://doi.org/10.1172/JCI116719.
Received in revised form 4 December 2017;
[9] I. Juhasz, G.F. Murphy, H.C. Yan, M. Herlyn, S.M. Albelda,
Accepted 1 January 2018 Regulation of extracellular matrix proteins and integrin cell
Available online 10 January 2018 substratum adhesion receptors on epithelium during cuta-
neous human wound healing in vivo, Am. J. Pathol. 143
Keywords: (1993) 1458–1469.
Wound healing; [10] D.T. Woodley, Reepithelialization, in: R.A.F. Clark (Ed.),
Re-epithelialization; The Molecular and Cellular Biology of Wound Repair, 2nd
Epidermal regeneration; ed.Plenum Press, New York 1996, pp. 339–354.
Keratinocyte; [11] R.A. Yates, L.B. Nanney, R.E. Gates, L.E. King, Epidermal
growth factor and related growth factors, Int. J. Dermatol. 30
Extracellular matrix;
(1991) 687–694, https://doi.org/10.1111/j.1365-4362.1991.
Basement membrane;
tb02609.x.
Laminin [12] G. Schultz, D.S. Rotatori, W. Clark, EGF and TGF-α in wound
healing and repair, J. Cell. Biochem. 45 (1991) 346–352.
[13] A. Michopoulou, P. Rousselle, How do epidermal matrix
metalloproteinases support re-epithelialization during skin
healing? Eur. J. Dermatol. 25 (2015) 33–42, https://doi.org/
10.1684/ejd.2015.2553.
Abbreviations used: [14] R.K. Sivamani, M.S. Garcia, R.R. Isseroff, Wound re-
3D, three-dimensional; BM, basement membrane; BMP-1, epithelialization: modulating keratinocyte migration in wound
bone morphogenetic protein-1; BPA, bullous pemphigoid healing, Front. Biosci. 12 (2007) 2849–2868, https://doi.org/
10.2741/2277.
antigen; DEJ, dermal-epidermal junction; EIII, extra type
[15] R.A.F. Clark, Wound repair: overview and general consid-
III homology; ECM, extracellular matrix; ED, extra domain; erations, in: R.A. Clark (Ed.), The Molecular and Cellular
EGF, epidermal growth factor; FN, fibronectin; GAG, Biology of Wound Repair, 2nd ed.Plenum Press, New York
glycosaminoglycan; HSPG, heparan sulfate proteoglycan; 1996, pp. 3–50.
IL, interleukin; LG, laminin globular domain; LM, laminin; [16] D. Breitkreutz, I. Koxholt, K. Thiemann, R. Nischt, Skin
MMPs, matrix metalloproteinases; mTLD, mammalian basement membrane: the foundation of epidermal integrity-
tolloid; OPN, osteopontin; RDEB, recessive dystrophic BM functions and diverse roles of bridging molecules nidogen
epidermolysis bullosa; SPARC, secreted protein acidic and perlecan, Biomed Res. Int. 2013 (2013) E-pub. Article ID
and rich in cysteine; TGF, transforming growth factor; 179784 https://doi.org/10.1155/2013/179784.
TSP, thrombospondin. [17] S. Werner, R. Grose, Regulation of wound healing by growth
factors and cytokines, Physiol. Rev. 83 (2003) 835–870,
https://doi.org/10.1152/physrev.00031.2002.
[18] J. Röhl, A. Zaharia, M. Rudolph, R.Z. Murray, The role of
References inflammation in cutaneous repair, Wound Pract. Res. 23
(2015) 8–15.
[19] R.A.F. Clark, J.M. Lanigan, P. DellaPelle, E. Manseau, H.F.
[1] R.A.F. Clark, Wound repair. Lessons for tissue engineering, in: Dvorak, R.B. Colvin, Fibronectin and fibrin provide a
R.P. Lanza, R. Langer, W. Chick (Eds.), Principles of Tissue provisional matrix for epidermal cell migration during wound
Engineering, Academic Press, San Diego 1997, pp. 737–768. reepithelialization, J. Invest. Dermatol. 79 (1982) 264–269.
[2] P. Martin, Wound healing-aiming for perfect skin regenera- [20] D. Chester, A.C. Brown, The role of biophysical properties
tion, Science 276 (1997) 75–81. of provisional matrix proteins in wound repair, Matrix Biol.
[3] T.J. Shaw, P. Martin, Wound repair at a glance, J. Cell Sci. 60–61 (2017) 124–140, https://doi.org/10.1016/j.matbio.
122 (2009) 3209–3213, https://doi.org/10.1242/jcs.031187. 2016.08.004.
[4] M.H. Kim, W. Liu, D.L. Borjesson, F.R. Curry, L.S. Miller, [21] W.S. To, K.S. Midwood, Plasma and cellular fibronectin:
A.L. Cheung, et al., Dynamics of neutrophil infiltration during distinct and independent functions during tissue repair,
cutaneous wound healing and infection using fluorescence Fibrogenesis Tissue Repair 4 (2011) 21, https://doi.org/10.
imaging, J. Invest. Dermatol. 128 (2008) 1812–1820, https:// 1186/1755-1536-4-21.
doi.org/10.1038/sj.jid.5701223. [22] R.O. Hynes, K.M. Yamada, Fibronectins: multifunctional
[5] A.J. Singer, R.A.F. Clark, Cutaneous wound healing, N. modular glycoproteins, J. Cell Biol. 95 (1982) 369–377,
Engl. J. Med. 341 (1999) 738–746. https://doi.org/10.1083/jcb.95.2.369.
[6] M.J. Redd, L. Cooper, W. Wood, B. Stramer, P. Martin, [23] R.A.F. Clark, Fibronectin in the skin, J. Invest. Dermatol. 81
Wound healing and inflammation: embryos reveal the way to (1983) 475–479.
perfect repair, Philos. Trans. R. Soc. Lond. Ser. B Biol. Sci. [24] C. Ffrench-Constant, L. Van de Water, H.F. Dvorak, R.O.
359 (2004) 777–784, https://doi.org/10.1098/rstb.2004.1466. Hynes, Reappearance of an embryonic pattern of fibronectin
[7] N.S. Greaves, S.A. Iqbal, M. Baguneid, A. Bayat, The role splicing during wound healing in the adult rat, J. Cell Biol. 109
of skin substitutes in the management of chronic (1989) 903–914.
22 Extracellular matrix contribution to skin wound re-epithelialization

[25] D.J. Donaldson, J.T. Mahan, Fibrinogen and fibronectin as phages differentially express fibronectin and its splice
substrates for epidermal cell migration during wound variants and the extracellular matrix protein βIG-H3,
closure, J. Cell Sci. 62 (1983) 117–127. Scand. J. Immunol. 53 (2001) 386–392.
[26] K. Igisu, The role of fibronectin in the process of wound [44] H. Larjava, L. Koivisto, L. Häkkinen, Keratinocyte interac-
healing, Thromb. Res. 44 (1986) 455–465. tions with fibronectin during wound healing, Madame Curie
[27] P. Carmeliet, L. Schoonjans, L. Kieckens, B. Ream, J. Degen, Bioscience Database [Internet], Austin (TX), Landes Bio-
R. Bronson, R. De Vos, et al., Physiological consequences of science, 2000–2013 Available from https://www.ncbi.nlm.
loss of plasminogen activator gene function in mice, Nature nih.gov/books/NBK6391/.
368 (1994) 419–424. [45] P. Singh, C.L. Reimer, J.H. Peters, M.A. Stepp, R.O. Hynes,
[28] T.H. Bugge, M.J. Flick, M.J.S. Danton, C.C. Daugherty, J. L. Van De Water, The spatial and temporal expression
Rømer, K. Danø, P. Carmeliet, et al., Urokinase-type patterns of integrin alpha9beta1 and one of its ligands, the
plasminogen activator is effective in fibrin clearance in the EIIIA segment of fibronectin, in cutaneous wound healing, J.
absence of its receptor or tissue-type plasminogen activa- Invest. Dermatol. 123 (2004) 1176–1181.
tor, Proc. Natl. Acad. Sci. 93 (1996) 5899–5904. [46] A.F. Muro, A.K. Chauhan, S. Gajovic, A. Iaconcig, F. Porro,
[29] J. Rømer, T.H. Bugge, C. Pyke, L.R. Lund, M.J. Flick, J.L. G. Stanta, et al., Regulated splicing of fibronectin EDA exon
Degen, et al., Impaired wound healing in mice with a is essential for proper skin wound healing and normal
disrupted plasminogen gene, Nat. Med. 2 (1996) 287–292. lifespan, J. Cell Biol. 162 (2003) 149–160.
[30] V. Ronfard, Y. Barrandon, Migration of keratinocytes through [47] T. Fukuda, N. Yoshida, Y. Kataoka, R. Manabe, Y. Mizuno-
tunnels of digested fibrin, Proc. Natl. Acad. Sci. U. S. A. 98 Horikawa, M. Sato, et al., Mice lacking the edb segment of
(2001) 4504–4509, https://doi.org/10.1073/pnas.071631698. fibronectin develop normally but exhibit reduced cell growth
[31] R.A.F. Clark, Fibronectin matrix deposition and fibronectin and fibronectin matrix assembly in vitro, Cancer Res. 62
receptor expression in healing and normal skin, J. Invest. (2002) 5603–5610.
Dermatol. 94 (1990) 1285–1345. [48] P. Singh, C. Carraher, J.E. Schwarzbauer, Assembly of
[32] A. Cavani, G. Zambruno, A. Marconi, V. Manca, M. fibronectin extracellular matrix, Annu. Rev. Cell Dev. Biol. 26
Marchetti, A. Giannetti, Distinctive integrin expression in (2010) 397–419, https://doi.org/10.1146/annurev-cellbio-
the newly forming epidermis during wound healing in 100109-104020.
humans, J. Invest. Dermatol. 101 (1993) 600–604. [49] J.E. Schwarzbauer, J.L. Sechler, Fibronectin fibrillogenesis:
[33] J. Gailit, M.P. Welch, R.A.F. Clark, TGF-beta 1 stimulates a paradigm for extracellular matrix assembly, Curr. Opin.
expression of keratinocyte integrins during re- Cell Biol. 11 (1999) 622–627.
epithelialization of cutaneous wounds, J. Invest. Dermatol. [50] J. Sottile, D.C. Hocking, Fibronectin polymerization regu-
103 (1994) 221–227. lates the composition and stability of extracellular matrix
[34] T. Sakai, K.J. Johnson, M. Murozono, K. Sakai, M.A. fibrils and cell-matrix adhesions, Mol. Biol. Cell. 13 (2002)
Magnuson, T. Wieloch, et al., Plasma fibronectin supports 3546–3559.
neuronal survival and reduces brain injury following [51] F. Shi, J. Harman, K. Fujiwara, J. Sottile, Collagen I matrix
transient focal cerebral ischemia but is not essential for turnover is regulated by fibronectin polymerization, Am. J.
skin-wound healing and hemostasis, Nat. Med. 7 (2001) Physiol. Cell Physiol. 298 (2009) 1265–1275.
324–330. [52] D.H. Madsen, D. Leonard, A. Masedunskas, A. Moyer, H.J.
[35] T.H. Barker, A.J. Engler, The provisional matrix: setting the Jürgensen, D.E. Peters, et al., M2-like macrophages are
stage for tissue repair outcomes, Matrix Biol. 60–61 (2017) responsible for collagen degradation through a mannose
1–4, https://doi.org/10.1016/j.matbio.2017.04.003. receptor-mediated pathway, J. Cell Biol. 202 (2013)
[36] F. Grinnell, R.E. Billingham, L. Burgess, Distribution of 951–966.
fibronectin during wound healing in vivo, J. Invest. [53] R. Raghow, The role of extracellular matrix in postinflam-
Dermatol. 76 (1981) 181–189. matory wound healing and fibrosis, FASEB J. 8 (1994)
[37] A.J. Zolinger, M.L. Smith, Fibronectin, the extracellular glue, 823–831.
Matrix Biol. 60–61 (2017) 27–37. [54] M.E. Nimmi, Collagen: its structure and function in normal
[38] Y. Mao, J.E. Schwarzbauer, Stimulatory effects of a three- and pathological connective tissues, Semin. Arthritis
dimensional microenvironment on cell-mediated fibronectin Rheum. 4 (1974) 95–150.
fibrillogenesis, J. Cell Sci. 118 (2005) 4427–4436. [55] J.N. Clore, I.K. Cohen, R.F. Diegelmann, Quantitation of
[39] J.E. Schwarzbauer, Identification of the fibronectin se- collagen types I and III during wound healing in rat skin,
quences required for assembly of a fibrillar matrix, J. Cell Exp. Biol. Med. 161 (1979) 337–340.
Biol. 113 (1991) 1463–1473. [56] M. Kurkinen, A. Vaheri, P.J. Roberts, S. Stenman, Sequential
[40] C. Ffrench-Constant, Alternative splicing of fibronectin- appearance of fibronectin and collagen in experimental
many different proteins but few different functions, Exp. granulation tissue, Lab. Investig. 43 (1980) 47–51.
Cell Res. 221 (1995) 261–271. [57] J.K. Mäkelä, E. Vuorio, Type I collagen messenger RNA
[41] A.R. Kornblihtt, C.G. Pesce, C.R. Alonso, P. Cramer, A. levels in experimental granulation tissue and silicosis in
Srebrow, S. Werbajh, et al., The fibronectin gene as a model rats, Med. Biol. 64 (1986) 15–22.
for splicing and transcription studies, FASEB J. 10 (1996) [58] R.A. Ignotz, J. Massaugue, Transforming growth factor beta
248–257. stimulates the expression of fibronectin and collagen and
[42] L.F. Brown, D. Dubin, L. Lavigne, B. Logan, H.F. Dvorak, L. their incorporation into the extracellular matrix, J. Biol.
Van de Water, Macrophages and fibroblasts express Chem. 261 (1986) 4337–4345.
embryonic fibronectins during cutaneous wound healing, [59] K. Scharffetter, M. Kulozik, W. Stolz, B. Lankat-Buttgereit,
Am. J. Pathol. 142 (1993) 793–801. A. Hatamochi, R. Söhnchen, et al., Localization of collagen
[43] A. Gratchev, P. Guillot, N. Hakiy, O. Politz, C.E. Orgfanos, alpha 1(I) gene expression during wound healing by in situ
K. Schledzewski, et al., Alternatively activated macro- hybridization, J. Invest. Dermatol. 93 (1989) 405–412.
Extracellular matrix contribution to skin wound re-epithelialization 23

[60] H.P. Ehrlich, T.M. Krummel, Regulation of wound healing macrophages: a new dimension in macrophage functional
from a connective tissue perspective, Wound Repair heterogeneity, J. Immunol. 180 (2008) 5707–5719.
Regen. 4 (1996) 203–210. [76] A. Nyström, D. Velati, V.R. Mittapalli, A. Fritsch, J.S.
[61] M. Guo, K. Toda, F. Grinnell, Activation of human Kern, L. Bruckner-Tuderman, Collagen VII plays a dual
keratinocyte migration on type I collagen and fibronectin, role in wound healing, J. Clin. Invest. 123 (2013)
J. Cell Sci. 96 (1990) 197–205. 3498–3509.
[62] M.J. Petersen, D.T. Woodley, G.P. Stricklin, E.J. O'Keefe, [77] D.T. Woodley, Y. Hou, S. Martin, W. Li, M. Chen,
Enhanced synthesis of collagenase by human keratinocytes Characterization of molecular mechanisms underlying
cultured on type I or type IV collagen, J. Invest. Dermatol. 94 mutations in dystrophic epidermolysis bullosa using site-
(1990) 341–346. directed mutagenesis, J. Biol. Chem. 283 (2008)
[63] B.K. Pilcher, J.A. Dumin, B.D. Sudbeck, S.M. Krane, H.G. 17838–17845.
Welgus, W.C. Parks, The activity of collagenase-1 is [78] X. Wang, P. Ghasri, M. Amir, B. Hwang, Y. Hou, M. Khalili,
required for keratinocyte migration on a type I collagen et al., Topical application of recombinant type VII collagen
matrix, J. Cell Biol. 137 (1997) 1445–1457. incorporates into the dermal-epidermal junction and pro-
[64] U.K. Saarialho-Kere, S.O. Kovacs, A.P. Pentland, J.E. motes wound closure, Mol. Ther. 21 (2013) 1335–1344.
Olerud, H.G. Welgus, W.C. Parks, Cell-matrix interactions [79] P. Betz, A. Nerlich, J. Wilske, J. Tübel, I. Wiest, R. Penning,
modulate interstitial collagenase expression by human et al., The time-dependent rearrangement of the epithelial
keratinocytes actively involved in wound healing, J. Clin. basement membrane in human skin wounds—immunohis-
Invest. 92 (1993) 2858–2866. tochemical localization of collagen IV and VII, Int. J. Legal
[65] D.T. Woodley, P.M. Bachmann, E.J. O'Keefe, Laminin Med. 105 (1992) 93–97.
inhibits human keratinocyte migration, J. Cell. Physiol. 136 [80] H. Qiao, A. Shibaki, H.A. Long, G. Wang, Q. Li, W. Nishie,
(1988) 140–146. et al., Collagen XVII participates in keratinocyte adhesion to
[66] M. Kubo, L. Van de Water, L.C. Plantefaber, M.W. Mosesson, collagen IV, and in p38MAPK-dependent migration and cell
M. Simon, M.G. Tonnesen, et al., Fibrinogen and fibrin are signalling, J. Invest. Dermatol. 129 (2009) 2288–2295.
anti-adhesive for keratinocytes: a mechanism for fibrin eschar [81] J. Jacków, A. Schlosser, R. Sormunen, A. Nyström, C.
slough during wound repair, J. Invest. Dermatol. 117 (2001) Sitaru, K. Tasanen, et al., Generation of a functional non-
1369–1381. shedding collagen xvii mouse model: relevance of collagen
[67] C. Woltersdorf, M. Bonk, B. Leitinger, M. Huhtala, J. Käpylä, XVII shedding in wound healing, J. Invest. Dermatol. 136
J. Heino, et al., The binding capacity of α1β1-, α2β1- and (2016) 516–525.
α10β1-integrins depends on non-collagenous surface mac- [82] J. Jacków, S. Löffek, A. Nyström, L. Bruckner-Tuderman,
romolecules rather than the collagens in cartilage fibrils, C.W. Franzke, Collagen XVII shedding suppresses re-
Matrix Biol. 63 (2017) 91–105. epithelialization by directing keratinocyte migration and
[68] A.D. Berendsen, A.L. Bronckers, T.H. Smit, X.F. dampening mtor signalling, J. Invest. Dermatol. 136
Walboomers, V. Everts, Collagen type V enhances matrix (2016) 1031–1041.
contraction by human periodontal ligament fibroblasts [83] S. Löffek, T. Hurskainen, J. Jacków, F.C. Sigloch, O.
seeded in three-dimensional collagen gels, Matrix Biol. 25 Schilling, K. Tasanen, et al., Transmembrane collagen XVII
(2006) 515–522. modulates integrin dependent keratinocyte migration via
[69] P. Martin, W.R. Teodoro, A.P. Velosa, J. de Morais, S. PI3K/Rac1 signaling, PLoS One 9 (2014), e87263. https://
Carrasco, R.B. Christmann, Abnormal collagen V deposi- doi.org/10.1371/journal.pone.0087263.
tion in dermis correlates with skin thickening and disease [84] L. Seppinen, R. Sormunen, Y. Soini, H. Elamaa, R.
activity in systemic sclerosis, Autoimmun. Rev. 11 (2012) Heljasvaara, T. Pihlajaniemi, Lack of collagen XVIII acceler-
827–835. ates cutaneous wound healing, while overexpression of its
[70] R.J. Wenstrup, J.B. Florer, J.M. Davidson, C.L. Phillips, B.J. endostatin domain leads to delayed healing, Matrix Biol. 27
Pfeiffer, D.W. Menezes, Murine model of the Ehlers–Danlos (2008) 535–546.
syndrome. col5a1 haploinsufficiency disrupts collagen fibril [85] H.T. Ren, H. Hu, Y. Li, H.F. Jiang, X.L. Hu, C.M. Han,
assembly at multiple stages, J. Biol. Chem. 281 (2006) Endostatin inhibits hypertrophic scarring in a rabbit ear
12888–12895. model, J Zhejiang Univ Sci B 14 (2013) 224–230.
[71] J. DeNigris, Q. Yao, E.K. Birk, D.E. Birk, Altered dermal [86] M. Aumailley, L. Bruckner-Tuderman, W.G. Carter, R.
fibroblast behavior in a collagen V haploinsufficient murine Deutzmann, D. Edgar, P. Ekblom, et al., A simplified
model of classic Ehlers–Danlos syndrome, Connect. Tissue laminin nomenclature, Matrix Biol. 24 (2005) 326–332.
Res. 57 (2016) 1–9. [87] M. Aumailley, P. Rousselle, Laminins of the dermo-
[72] P. Betz, A. Nerlich, J. Tubel, R. Penning, W. Eisenmenger, epidermal junction, Matrix Biol. 18 (1999) 19–28.
Localization of tenascin in human skin wounds-an immuno- [88] M.C. Ryan, R. Tizard, D.R. VanDevanter, W.G. Carter,
histochemical study, Int. J. Legal Med. 105 (1993) 325–328. Cloning of the LamA3 gene encoding the alpha 3 chain of
[73] T. Oono, U. Specks, B. Eckes, S. Majewski, N. the adhesive ligand epiligrin. Expression in wound repair, J.
Hunzelmann, R. Timpl, et al., Expression of type VI Biol. Chem. 269 (1994) 22779–22787.
collagen mRNA during wound healing, J. Invest. Dermatol. [89] M.C. Ryan, K. Lee, Y. Miyashita, W.G. Carter, Targeted
100 (1993) 329–334. disruption of the LAMA3 gene in mice reveals abnormalities
[74] G. Theocharidis, Z. Drymoussi, A.P. Kao, A.H. Barber, D.A. in survival and late stage differentiation of epithelial cells, J.
Lee, K.M. Braun, et al., Type VI collagen regulates dermal Cell Biol. 145 (1999) 1309–1323.
matrix assembly and fibroblast motility, J. Invest. Dermatol. [90] M.A. Kurpakus, E.L. Stock, J.C. Jones, Analysis of wound
136 (2016) 74–83. healing in an in vitro model: early appearance of laminin and
[75] M. Schnoor, P. Cullen, J. Lorkowski, K. Stolle, H. Robenek, a 125 × 103 Mr polypeptide during adhesion complex
D. Troyer, et al., Production of type VI collagen by human formation, J. Cell Sci. 96 (1990) 651–660.
24 Extracellular matrix contribution to skin wound re-epithelialization

[91] B.P. Nguyen, S.G. Gil, W.G. Carter, Deposition of laminin 5 [108] R.O. Sigle, S.G. Gil, M. Bhattacharya, M.C. Ryan, T.M.
by keratinocytes regulates integrin adhesion and signalling, Yang, T.A. Brown, et al., Globular domains 4/5 of the
J. Biol. Chem. 275 (2000) 31896–31907. laminin α3 chain mediate deposition of precursor laminin 5,
[92] A.F. Laplante, L. Germain, F.A. Auger, V. Moulin, Mecha- J. Cell Sci. 117 (2004) 4481–4494.
nisms of wound reepithelialization: hints from a tissue- [109] L.E. Goldfinger, M.S. Stack, J.C. Jones, Processing of
engineered reconstructed skin to long-standing questions, laminin-5 and its functional consequences: role of plasmin
FASEB J. 15 (2001) 2377–2389. and tissue-type plasminogen activator, J. Cell Biol. 141
[93] M. Aragona, S. Dekoninck, S. Rulands, S. Lenglez, G. Mascré, (1998) 255–265.
B.D. Simons, et al., Defining stem cell dynamics and migration [110] O. Okamoto, S. Bachy, U. Odenthal, J. Bernaud, D. Rigal,
during wound healing in mouse skin epidermis, Nat. Commun. H. Lortat-Jacob, et al., Normal human keratinocytes bind to
8 (2017) 1–14, https://doi.org/10.1038/ncomms14684. the α3LG4/5 domain of unprocessed laminin-5 through the
[94] P. Rousselle, K. Beck, Laminin 332 processing impacts receptor syndecan-1, J. Biol. Chem. 278 (2003) 44168–44177.
cellular behaviour, Cell Adhes. Migr. 7 (2013) 122–134, [111] A. Utani, Y. Momota, H. Endo, Y. Kasuya, K. Beck, N.
https://doi.org/10.4161/cam.23132. Suzuki, et al., Laminin α3 LG4 module induces matrix
[95] L. Tunggal, J. Ravaux, M. Pesch, H. Smola, T. Krieg, F. metalloproteinase-1 through mitogen-activated protein ki-
Gaill, et al., Defective laminin 5 processing in cylindroma nase signalling, J. Biol. Chem. 278 (2003) 34483–34490.
cells, Am. J. Pathol. 160 (2002) 459–468. [112] S. Carulli, K. Beck, G. Dayan, S. Boulesteix, H. Lortat-Jacob,
[96] M. Aumailley, A. El Khal, N. Knöss, L. Tunggal, Laminin 5 P. Rousselle, Cell surface proteoglycans syndecan-1 and -4
processing and its integration into the ECM, Matrix Biol. 22 bind overlapping but distinct sites in laminin α3 LG45 protein
(2003) 49–54. domain, J. Biol. Chem. 287 (2012) 12204–12216.
[97] P. Rousselle, G.P. Lunstrum, D.R. Keene, R.E. Burgeson, [113] B. Sulka, H. Lortat-Jacob, R. Terreux, F. Letourneur, P.
Kalinin: an epithelium-specific basement membrane adhe- Rousselle, Tyrosine dephosphorylation of the syndecan-1
sion molecule that is a component of anchoring filaments, J. PDZ binding domain regulates syntenin-1 recruitment, J.
Cell Biol. 114 (1991) 567–576. Biol. Chem. 284 (2009) 10659–10671.
[98] W.G. Carter, M.C. Ryan, P.J. Gahr, Epiligrin, a new cell [114] I. Senyürek, W.E. Kempf, G. Klein, A. Maurer, H. Kalbacher,
adhesion ligand for integrin α3β1 in epithelial basement L. Schäfer, et al., Processing of laminin α chains generates
membranes, Cell 65 (1991) 599–610. peptides involved in wound healing and host defense, J.
[99] A. Sonnenberg, A.A. de Melker, A.M. Martinez de Velasco, H. Innate. Immun. 6 (2014) 467–484.
Janssen, J. Calafat, C.M. Niessen, Formation of hemidesmo- [115] E. Natarajan, M. Saeb, C.P. Crum, S.B. Woo, P.H. McKee,
somes in cells of a transformed murine mammary tumor cell line J.G. Rheinwald, Co-expression of p16(INK4A) and laminin
and mechanisms involved in adherence of these cells to laminin 5 gamma2 by microinvasive and superficial squamous cell
and kalinin, J. Cell Sci. 106 (1993) 1083–1102. carcinomas in vivo and by migrating wound and senescent
[100] P. Rousselle, M. Aumailley, Kalinin is more efficient than keratinocytes in culture, Am. J. Pathol. 163 (2003) 477–491.
laminin in promoting adhesion of primary keratinocytes and [116] E. Natarajan, J.D. Omobono 2nd, Z. Guo, S. Hopkinson,
some other epithelial cells and has a different requirement A.J. Lazar, T. Brenn, et al., A keratinocyte hypermotility/
for integrin receptors, J. Cell Biol. 125 (1994) 205–214. growth-arrest response involving laminin 5 and p16INK4a
[101] H. Mizushima, H. Takamura, Y. Miyagi, Y. Kikkawa, N. activated in wound healing and senescence, Am. J. Pathol.
Yamanaka, H. Yasumitsu, et al., Identification of integrin- 168 (2006) 1821–1837.
dependent and -independent cell adhesion domains in [117] L. Gagnoux-Palacios, M. Allegra, F. Spirito, O. Pommeret,
COOH-terminal globular region of laminin-5 α3 chain, Cell C. Romero, J.P. Ortonne, et al., The short arm of the laminin
Growth Differ. 8 (1997) 979–987. gamma2 chain plays a pivotal role in the incorporation of
[102] T. Kainulainen, L. Hakkinen, S. Hamidi, K. Larjava, M. laminin 5 into the extracellular matrix and in cell adhesion, J.
Kallioinen, J. Peltonen, et al., Laminin-5 expression is Cell Biol. 153 (2001) 835–850.
independent of the injury and the microenvironment during [118] G. Fisher, L. Rittié, Restoration of the basement membrane
reepithelialization of wounds, J. Histochem. Cytochem. 46 after wounding: a hallmark of young human skin altered with
(1998) 353–360. aging, J. Cell Commun. Signal. (2017)https://doi.org/10.
[103] P.D. Lampe, B.P. Nguyen, S. Gil, M. Usui, J. Olerud, Y. 1007/s12079-017-0417-3.
Takada, et al., Cellular interaction of integrin alpha3beta1 [119] W.M. Longmate, R. Monichan, M.L. Chu, T. Tsuda, M.G.
with laminin 5 promotes gap junctional communication, J. Mahoney, C.M. DiPersio, Reduced fibulin-2 contributes to
Cell Biol. 143 (1998) 1735–1747. loss of basement membrane integrity and skin blistering in
[104] L.E. Goldfinger, S.B. Hopkinson, G.W. deHart, S. Collawn, mice lacking integrin alpha3beta1 in the epidermis, J.
J.R. Couchman, J.C. Jones, The α3 laminin subunit, α6β4 Invest. Dermatol. 134 (2014) 1609–1617.
and α3β1 integrin coordinately regulate wound healing in [120] W.M. Longmate, S.P. Lyons, L. DeFreest, L. Van De Water,
cultured epithelial cells and in the skin, J. Cell Sci. 112 C.M. DiPersio, Opposing roles of epidermal integrins α3β1
(1999) 2615–2629. and α9β1 in regulation of mTLD/BMP-1-mediated laminin-
[105] F. Decline, P. Rousselle, Keratinocyte migration requires γ2 processing during wound healing, J. Invest. Dermatol.
α2β1 integrin-mediated interaction with the laminin 5 γ2 (2017) 32945–32947, https://doi.org/10.1016/j.jid.2017.09.
chain, J. Cell Sci. 114 (2001) 811–823. 004.
[106] E. Hintermann, V. Quaranta, Epithelial cell motility on [121] A.M. Muir, D. Massoudi, N. Nguyen, D.R. Keene, S.J. Lee,
laminin-5: regulation by matrix assembly, proteolysis, D.E. Birk, et al., BMP1-like proteinases are essential to the
integrins and erbB receptors, Matrix Biol. 23 (2004) 75–85. structure and wound healing of skin, Matrix Biol. 56 (2016)
[107] D.E. Frank, W.G. Carter, Laminin 5 deposition regulates 114–131.
keratinocyte polarization and persistent migration, J. Cell [122] N. Pouliot, N.A. Saunders, P. Kaur, Laminin 10/11: an
Sci. 117 (2004) 1351–1363. alternative adhesive ligand for epidermal keratinocytes with
Extracellular matrix contribution to skin wound re-epithelialization 25

a functional role in promoting proliferation and migration, [139] E. Kohfeldt, T. Sasaki, W. Göhring, R. Timpl, Nidogen-2: a
Exp. Dermatol. 11 (2002) 387–397. new basement membrane protein with diverse binding
[123] A. Botta, F. Delteil, A. Mettouchi, A. Viera, S. Estrach, L. properties, J. Mol. Biol. 282 (1998) 99–109.
Négroni, et al., Confluence switch signaling regulates ECM [140] B.L. Bader, N. Smyth, S. Nedbal, N. Miosge, A. Baranowsky,
composition and plasmin proteolytic cascade in keratino- S. Mokkapati, et al., Compound genetic ablation of nidogen 1
cytes, J. Cell Sci. 125 (2012) 4341–4352. and 2 causes basement membrane defects and perinatal
[124] C. Has, A. Nyström, Epidermal basement membrane in lethality in mice, Mol. Cell. Biol. 25 (2005) 6846–6856.
health and disease. In basement membranes, Curr. Top. [141] S. Mokkapati, A. Baranowsky, N. Mirancea, N. Smyth, D.
Membr. 76 (2015) 117–170. Breitkreutz, R. Nischt, Basement membranes in skin are
[125] L.M. Sorokin, F. Pausch, M. Frieser, S. Kroger, E. Ohage, R. differently affected by lack of nidogen 1 and 2, J. Invest.
Deutzmann, Developmental regulation of the laminin alpha5 Dermatol. 128 (2008) 2259–2267.
chain suggests a role in epithelial and endothelial cell [142] A. Baranowsky, S. Mokkapati, M. Bechtel, J. Krügel, N.
maturation, Dev. Biol. 189 (1997) 285–300. Miosge, C. Wickenhauser, et al., Impaired wound healing in
[126] S. Paquet-Fifield, H. Schlüter, A. Li, T. Aitken, P. mice lacking the basement membrane protein nidogen 1,
Gangatirkar, D. Blashk, et al., A role for pericytes as Matrix Biol. 29 (2010) 15–21.
microenvironmental regulators of human skin tissue regen- [143] P. Bornstein, E.H. Sage, Matricellular proteins: extracellular
eration, J. Clin. Invest. 119 (2009) 2795–2806. modulators of cell function, Curr. Opin. Cell Biol. 14 (2002)
[127] J. Wegner, K. Loser, G. Apsite, R. Nischt, B. Eckes, T. 608–616.
Krieg, et al., Laminin α5 in the keratinocyte basement [144] T.R. Kyriakides, S. MacLauchlan, The role of thrombos-
membrane is required for epidermal-dermal intercommuni- pondins in wound healing, ischemia, and the foreign body
cation, Matrix Biol. 56 (2016) 24–41. reaction, J. Cell Commun. Signal. 3 (2009) 215–225.
[128] R. Timpl, J.C. Brown, Supramolecular assembly of base- [145] J.E. Murphy-Ullrich, M. Poczatek, Activation of latent TGF-
ment membranes, BioEssays 18 (1996) 123–132. b1 by thrombospondin-1: mechanisms and physiology,
[129] D.T. Behrens, D. Villone, M. Koch, G. Brunner, L. Sorokin, Cytokine Growth Factor Rev. 11 (2000) 59–69.
H. Robenek, et al., The epidermal basement membrane is a [146] S.E. Crawford, V. Stellmach, J.E. Murphy-Ullrich, S.M.
composite of separate laminin- or collagen IV-containing Ribeiro, J. Lawler, R.O. Hynes, et al., Thrombospondin-1 is
networks connected by aggregated perlecan, but not by a major activator of TGF-beta1 in vivo, Cell 93 (1998)
nidogens, J. Biol. Chem. 22 (2012) 18700–18709. 1159–1170.
[130] P.D. Yurchenco, Integrating activities of laminins that drive [147] T.R. Kyriakides, J.W. Tam, P. Bornstein, Accelerated
basement membrane assembly and function, Curr. Top. wound healing in mice with a disruption of the thrombos-
Membr. 76 (2015) 1–30. pondin 2 gene, J. Invest. Dermatol. 113 (1999) 782–787.
[131] J.R. Hassel, P.G. Robey, H.J. Barrach, J. Wilczek, S.I. [148] A. Agah, T.R. Kyriakides, J. Lawler, P. Bornstein, The lack
Rennard, G.R. Martin, Isolation of a heparan sulfate- of thrombospondin-1 (TSP1) dictates the course of wound
containing proteoglycan from basement membrane, Proc. healing in double-TSP1/TSP2-null mice, Am. J. Pathol. 161
Natl. Acad. Sci. U. S. A. 77 (1980) 4494–4498. (2002) 831–839.
[132] M. Paulsson, S. Fujiwara, M. Dziadek, R. Timpl, G. Pejler, [149] L. DiPietro, N. Nissen, R. Gamelli, A. Koch, J. Pyle, P.
G. Backstrom, et al., Structure and function of basement Polverini, Thrombospondin 1 synthesis and function in
membrane proteoglycans, CIBA Found. Symp. 124 (1986) wound repair, Am. J. Pathol. 148 (1996) 1851–1860.
189–203. [150] T.R. Kyriakides, Y.H. Zhu, Z. Yang, G. Huynh, P. Bornstein,
[133] R.V. Iozzo, J.J. Zoeller, A. Nyström, Basement membrane Altered extracellular matrix remodeling and angiogenesis in
proteoglycans: modulators par excellence of cancer growth sponge granulomas of thrombospondin 2-null mice, Am. J.
and angiogenesis, Mol Cells. 27 (2009) 503–513. Pathol. 159 (2001) 1255–1262.
[134] Z. Zhou, J. Wang, R. Cao, H. Morita, R. Soininen, K.M. [151] S. MacLauchlan, E.A. Skokos, A. Agah, J. Zeng, W. Tian, J.M.
Chan, et al., Impaired angiogenesis, delayed wound healing Davidson, et al., Enhanced angiogenesis and reduced
and retarded tumor growth in perlecan heparan sulfate- contraction in thrombospondin-2-null wounds is associated
deficient mice, Cancer Res. 64 (2004) 4699–4702. with increased levels of matrix metalloproteinases-2 and -9, and
[135] I. Sher, S. Zisman-Rozen, L. Eliahu, J.M. Whitelock, E. soluble VEGF, J. Histochem. Cytochem. 57 (2009) 301–313.
Hirasawa, R.V. Iozzo, et al., Targeting perlecan in [152] D.R. Senger, D.F. Wirth, R.O. Hynes, Transformed mam-
human keratinocytes reveals novel roles for perlecan in malian cells secrete specific proteins and phosphoproteins,
epidermal formation, J. Biol. Chem. 281 (2006) Cell 16 (1979) 885–893.
5178–5187. [153] F. Buback, A.C. Renkl, G. Schulz, J.M. Weiss, Osteopontin
[136] M. Dos Santos, A. Michopoulou, V. André-Frei, S. and the skin: multiple emerging roles in cutaneous biology
Boulesteix, C. Guicher, G. Dayan, et al., Perlecan and pathology, Exp. Dermatol. 18 (2009) 750–759.
expression influences the keratin 15-positive cell popula- [154] L. Liaw, D.E. Birk, C.B. Ballas, J.S. Whitsitt, J.M. Davidson,
tion fate in the epidermis of aging skin, Aging 8 (2016) B.L. Hogan, Altered wound healing in mice lacking a
751–768. functional osteopontin gene (spp1), J. Clin. Invest. 101
[137] M. Aumailley, C. Battaglia, U. Mayer, R. Nischt, R. Timpl, (1998) 1468–1478.
J.W. Fox, Nidogen mediates the formation of ternary [155] R. Mori, T.J. Shaw, P. Martin, Molecular mechanisms linking
complexes of basement membrane components, Kidney wound inflammation and fibrosis: knockdown of osteopontin
Int. 43 (1993) 7–12. leads to rapid repair and reduced scarring, J. Exp. Med. 205
[138] J.W. Fox, U. Mayer, R. Nischt, M. Aumailley, D. Reinhardt, (2008) 43–51.
H. Wiedemann, et al., Recombinant nidogen consists of [156] K.S. Midwood, G. Orend, The role of tenascin-C in tissue
three globular domains and mediates binding of laminin to injury and tumorigenesis, J. Cell Commun. Signal. 3 (2009)
collagen IV, EMBO J. 10 (1991) 3137–3146. 287–310.
26 Extracellular matrix contribution to skin wound re-epithelialization

[157] M.A. Latijnhouwers, M. Bergers, B.H. Van Bergen, K.I. and thrombospondin 1 in wound repair: immunolocalization
Spruijt, M.P. Andriessen, J. Schalkwijk, Tenascin expres- and in situ hybridization, J. Histochem. Cytochem. 41
sion during wound healing in human skin, J. Pathol. 178 (1993) 1467–1477.
(1996) 30–35. [172] N. Hunzelman, M. Hafner, S. Anders, T. Krieg, R. Nischt,
[158] K.S. Midwood, M. Chiquet, R.P. Tucker, G. Orend, BM-40 (osteonectin, SPARC) is expressed both in the
Tenascin-C at a glance, J. Cell Sci. 129 (2016) 4321–4327. epidermal and in the dermal compartment of adult skin, J.
[159] R. Chiquet-Ehrismann, M. Chiquet, Tenascins: regulation Invest. Dermatol. 110 (1998) 122–126.
and putative functions during pathological stress, J. Pathol. [173] A.D. Bradshaw, M.J. Reed, E.H. Sage, SPARC-null mice
200 (2003) 488–499. exhibit accelerated cutaneous wound closure, J. Histo-
[160] I.A. Udalova, M. Ruhmann, S.J. Thomson, K.S. Midwood, chem. Cytochem. 50 (2002) 1–10.
Expression and immune function of tenascin-C, Crit. Rev. [174] L.E. Tracy, R.A. Minasian, E.J. Caterson, Extracellular
Immunol. 31 (2011) 115–145. matrix and dermal fibroblast function in the healing wound,
[161] M.M. Seyger, J.P. van Pelt, J. van den Born, M.A. Adv. Wound Care (New Rochelle). 5 (2016) 119–136.
Latijnhouwers, E.M. de Jong, Epicutaneous application of [175] S.A. Eming, T.A. Wynn, P. Martin, Inflammation and
leukotriene B4 induces patterns of tenascin and a heparan metabolism in tissue repair and regeneration, Science 356
sulfate proteoglycan epitope that are typical for psoriatic (2017) 1026–1030.
lesions, Arch. Dermatol. Res. 289 (1997) 331–336. [176] T. Lucas, A. Waisman, R. Ranjan, J. Roes, T. Krieg, W.
[162] R.A.C. Clark, H.P. Erickson, T.A. Springer, Tenascin supports Müller, et al., Differential roles of macrophages in diverse
lymphocyte rolling, J. Cell Biol. 137 (1997) 755–765. phases of skin repair, J. Immunol. 184 (2010) 3964–3977.
[163] H. Nakahara, E.C. Gabazza, H. Fujimoto, Y. Nishii, C.N. [177] P.A. Jones, T. Scott-Burden, Activated macrophages digest
D'Alessandro-Gabazza, N.E. Bruno, et al., Deficiency of the extracellular matrix proteins produced by cultured cells,
tenascin C attenuates allergen-induced bronchial asthma in Biochem. Biophys. Res. Commun. 86 (1979) 71–77.
the mouse, Eur. J. Immunol. 36 (2006) 3334–3345. [178] P.J. Murray, T.A. Wynn, Protective and pathogenic func-
[164] K. Midwood, S. Sacre, A.M. Piccinini, J. Inglis, A. Trebaul, E. tions of macrophage subsets, Nat. Rev. Immunol. 11 (2011)
Chan, et al., Tenascin-C is an endogenous activator of toll- 723–737.
like receptor 4 that is essential for maintaining inflammation [179] R. Richardson, K. Slanchev, C. Kraus, P. Knyphausen, S.
in arthritic joint disease, Nat. Med. 15 (2009) 774–780. Eming, M. Hammerschmidt, Adult zebrafish as a model
[165] E.J. Mackie, W. Halfter, D. Liverani, Induction of tenascin in system for cutaneous wound-healing research, J. Invest.
healing wounds, J. Cell Biol. 107 (1998) 2757–2767. Dermatol. 133 (2013) 1655–1665.
[166] M. Latijnhouwers, M. Bergers, M. Ponec, H. Dijkman, M. [180] J. Caddy, T. WIlanowski, C. Darido, S. Dworkin, S.B. Ting,
Andriessen, J. Schalkwijk, Human epidermal keratinocytes Q. Zhao, et al., Epidermal wound repair is regulated by the
are a source of tenascin-C during wound healing, J. Invest. planar cell polarity signaling pathway, Dev. Cell 19 (2010)
Dermatol. 108 (1997) 776–783. 138–147.
[167] R. Fassler, T. Sasaki, R. Timpl, M.L. Chu, S. Werner, Differential [181] R. Richardson, M. Metzger, P. Knyphausen, T. Ramezani,
regulation of fibulin, tenascin-C, and nidogen expression during K. Slanchev, C. Kraus, et al., Re-epithelialization of
wound healing of normal and glucocorticoid-treated mice, Exp. cutaneous wounds in adult zebrafish combines mecha-
Cell Res. 222 (1996) 111–116. nisms of wound closure in embryonic and adult mammals,
[168] A. Dalkowski, D. Schuppan, C.E. Orfanos, C.C. Zouboulis, Development 143 (2016) 2077–2088.
Increased expression of tenascin C by keloids in vivo and in [182] B. Garcia-Fernandez, I. Campos, J. Geiger, A.C. Santos, A.
vitro, Br. J. Dermatol. 141 (1999) 50–56. Jacinto, Epithelial resealing, Int. J. Dev. Biol. 53 (2009)
[169] W. Filsell, S. Rudman, G. Jenkins, M.R. Green, Coordinate 1549–1556.
upregulation of tenascin C expression with degree of photo- [183] R.O. Hynes, The dynamic dialogue between cells and
damage in human skin, Br. J. Dermatol. 140 (1999) 592–599. matrices: implications of fibronectin's elasticity, Proc. Natl.
[170] A.D. Bradshaw, E.H. Sage, SPARC, a matricellular protein Acad. Sci. U. S. A. 96 (1999) 2588–2590.
that functions in cellular differentiation and tissue response [184] J.E. Schwarzbauer, Alternative splicing of fibronectin: three
to injury, J. Clin. Invest. 107 (2001) 1049–1054. variants, three functions, BioEssays 13 (1991) 527–533.
[171] M.J. Reed, P. Puolakkainen, T.F. Lane, D. Dickerson, P.
Bornstein, E.H. Sage, Differential expression of SPARC

You might also like