Next Article in Journal
Role of Oxidative Stress in Drug-Induced Kidney Injury
Next Article in Special Issue
The Multifaceted Roles of PI3Kγ in Hypertension, Vascular Biology, and Inflammation
Previous Article in Journal
Assessment of Photodynamic Inactivation against Periodontal Bacteria Mediated by a Chitosan Hydrogel in a 3D Gingival Model
Previous Article in Special Issue
Progressive Vascular Functional and Structural Damage in a Bronchopulmonary Dysplasia Model in Preterm Rabbits Exposed to Hyperoxia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Extracellular Matrix, a Hard Player in Angiogenesis

Experimental Oncology Division 2, Department of Translational Research, CRO-IRCCS, Aviano 33081, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2016, 17(11), 1822; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17111822
Submission received: 22 August 2016 / Revised: 30 September 2016 / Accepted: 21 October 2016 / Published: 1 November 2016
(This article belongs to the Special Issue Vascular Biology and Therapeutics)

Abstract

:
The extracellular matrix (ECM) is a complex network of proteins, glycoproteins, proteoglycans, and polysaccharides. Through multiple interactions with each other and the cell surface receptors, not only the ECM determines the physical and mechanical properties of the tissues, but also profoundly influences cell behavior and many physiological and pathological processes. One of the functions that have been extensively explored is its impingement on angiogenesis. The strong impact of the ECM in this context is both direct and indirect by virtue of its ability to interact and/or store several growth factors and cytokines. The aim of this review is to provide some examples of the complex molecular mechanisms that are elicited by these molecules in promoting or weakening the angiogenic processes. The scenario is intricate, since matrix remodeling often generates fragments displaying opposite effects compared to those exerted by the whole molecules. Thus, the balance will tilt towards angiogenesis or angiostasis depending on the relative expression of pro- or anti-angiogenetic molecules/fragments composing the matrix of a given tissue. One of the vital aspects of this field of research is that, for its endogenous nature, the ECM can be viewed as a reservoir to draw from for the development of new more efficacious therapies to treat angiogenesis-dependent pathologies.

Graphical Abstract

1. Introduction

The extracellular matrix (ECM) is composed by a variety of proteins, glycoproteins, proteoglycans, and polysaccharides that are endowed with distinct physical and biochemical properties [1]. For long the ECM has been viewed as a scaffold with mere mechanical properties aimed at maintaining tissue morphology. A growing amount of evidences support instead the notion that ECM is surprisingly versatile and dynamic and can profoundly affect cell behavior [2]. In fact, through direct or indirect mechanisms the ECM modulates the function of the adjacent cells during development but also in the adulthood under normal as well as pathological conditions [3]. For its crucial role in assuring proper tissue homeostasis and function, the deposition and remodeling of the ECM components are tightly regulated [4]. If this regulation is lost, the matrix loses its organization and this leads to an impaired cell behavior and to a consequent break down of tissue function and homeostasis. In fact, abnormal ECM is one of the most distinctive traits that can be observed in many pathologies, including fibrosis and cancer [5]. Cancer is one of the major causes of mortality accounting for millions of deaths worldwide. Several ECM components have been shown to play a critical role, not only during its onset but also during metastasis, which is responsible for the 90% of all the cancer-related deaths [6,7,8,9,10]. In this context, the activation of proteases leads to the release of matrix fragments that act directly on cancer cells influencing their viability, apoptotic rate and/or metastatic potential [11,12,13,14,15,16]. Matrix remodeling leads also to the release of growth factors and cytokines of which the ECM represent a vital reservoir. The thus altered tumor microenvironment not only directly affects tumor cell behavior, but also other important processes such as angiogenesis. Angiogenesis, the process of the formation of new blood vessel from pre-existing vasculature, plays an indispensable role both in physiological and pathological conditions. It occurs during development and maturation, as well as during the healing of injured tissues and reproduction in the female population and is indispensable to grant the supply of nutrients and oxygen to the tissues. The angiogenic process is also usurped by a number of pathologies including inflammation, autoimmune diseases and cancer [17]. Notably an imbalanced vascularization can contribute to the development of these diseases [18]. In autoimmune diseases an excessive angiogenesis can promote inflammatory responses and, as a consequence, exacerbate the disorder. In cancer, angiogenesis is required to allow and efficient growth of the restlessly proliferating tumor cells. The imbalanced angiogenic signals released during tumor onset and progression lead to the development of aberrantly tortuous and leaky vessels. The altered vasculature not only represents an important root for metastatic spreading, but also affects the delivery of chemotherapy drugs [19,20]. Along these lines, a recent view about the therapies targeting angiogenesis is that they should induce the formation of a proficient vasculature that would favor the distribution of the chemotherapy drugs within the tumors. Angiogenesis can be not only an accomplice of the disease, but also be its direct cause as demonstrated for macular degeneration and diabetic retinopathy [21].
For its primary role in tissue trophism and homeostasis, the angiogenic process is tightly regulated by a plethora of factors. In fact, the list of molecules affecting angiogenesis is growing and includes growth factors, bioactive lipids, complex polysaccharides and also several ECM components. The most important cell type activated during the angiogenic process is the Endothelial cell (EC). These cells form the inner cell lining of blood vessels and are present as a single-layered epithelium adjacent to the lumen of the vessels. ECs are normally quiescent and characterized by a very low proliferative rate. In part this is due to the strong anchorage of these cells, to the tight cell junctions that grant structural continuity and limit vascular leak. The release of growth factors and the onset of strong pro-angiogenic stimuli loosen these restrictions. As a consequence, ECs change their behavior and begin to proliferate and migrate invading the surrounding tissues. This is followed by a resolution state where the ECs re-acquire the quiescent state. Despite a gradient of cytokines or other agonists is necessary to prompt EC migration, the movement of these cells is strictly dependent on their adhesion to the ECM and also for this reason these constituents play a key role in this process [22,23,24]. The number of ECM molecules that have been shown to influence angiogenesis is increasing and this depends not only on their adhesive properties but also due to their intrinsic capacity to directly affect EC function. Given the variety of ECM constituents, to study the contribution of the microenvironment in this context is extremely complex. To exacerbate the scenario, the fragmentation of ECM molecules occurring during remodeling leads to the formation of peptides that can often display opposite effects compared to the intact molecules of origin [25].
The aim of this article is to review a number of ECM molecules displaying a prominent role during angiogenesis and to summarize the major mechanisms impinging on EC function thus affecting the formation of new blood vessels. A schematic representation of the major classes of ECM molecules involved in angiogenesis is reported in Figure 1 and the molecules/fragments taken into account along with their respective functions summarized in Table A1.

2. Thrombospondins: Endogenous Angiogenesis Inhibitors

Thrombospondins (TSPs) are a family of large ECM glycoproteins including five members (TSP-1 to TSP-5) [26]. TSP-1 and TSP-2 have been extensively studied for their antiangiogenic properties. TSP-1 was the first to be identified as a naturally occurring angiogenic inhibitor [27]. Since this seminal paper, numerous have been the studies indicating that both TSP-1 and TSP-2 inhibit angiogenesis in multiple in vitro and in vivo assays [28,29]. Accordingly, the over-expression of these two molecules by tumor cells is associated with an impaired tumor growth in mice [30,31,32,33]. Interestingly, the simultaneous over-expression of both the molecules leads to additive effects, suggesting that TSP-1 and TSP-2 act by distinct mechanisms [31]. TSP-1 and TSP-2 specifically induce apoptosis in microvascular ECs leading to an impaired EC function and prejudiced tubule formation in vitro [34,35]. The molecular mechanism involves the binding to the transmembrane glycoprotein CD36 [34,36,37]. The interaction with the receptor results in the association of Src family kinases fyn or yes, which in turn lead to apoptosis through the phosphorylation of c-jun N-terminal kinase (JNK) and caspases [38]. It is interesting to note that the ECM non only supports cell survival through integrins engagement, but can also negatively affect cell viability in some circumstances.
In order to pinpoint the region(s) of TSP-1 responsible for the anti-angiogenic activity, synthetic peptides based on the protein sequence have been created and employed [28,35,39]. One of the first synthetic peptide generated contains the CSVTCG amino acidic sequence and was active in inhibiting the FGF-2-induced neovascularization of the rat cornea [28]. Some of the active sequences mapped are common to both TSP-1 and TSP-2, which indicates that they share common mechanisms of action. TSPs determine the EC phenotype through the binding to cell membrane or cell-associated molecules including integrins, CD36, CD47 and proteoglycans [40]. The multiple interactions on the EC surface may result in the assembly of molecular complexes able to trigger several signal transduction pathways, which may be the key of their strong effects. TSP-1 interacts and modulates the function of several cytokines/growth factors, proteases and ECM molecules [41], which suggests the possibility to develop new anti-angiogenic drugs [39,42,43]. Both TSP-1 and TSP-2 contain the thrombospondin type 1 repeat (TSR), an essential module for the binding to β1 integrins, to CD36 and to the transforming growth factor β (TGF-β) [44]. Notably, a therapeutic agent based on the second TSR has been developed [44] and is in clinical trials for several malignancies [45,46].
Thus, the studies on TSP-1 and TSP-2 have lead to important findings that have the translational potential to generate new promising tools for clinical practice. The drugs based on endogenous antiangiogenic molecules may in fact enclose the advantage not to induce harsh side effects during the treatments.

3. Fibronectin: Key Function in Pathological Angiogenesis

Fibronectin is strongly expressed around developing vessels during embryogenesis but is barely detectable in the adult vasculature [47]. Nonetheless, during pathological angiogenesis, the expression of the molecule is turned back on [48,49]. Fibronectin forms dimers composed of two similar monomers, which are organized into type I, type II and type III repeats. Different isoforms are present and are generated by alternative splicing within the EIIIA, EIIIB and V regions.
Fibronectin interacts with different ECM components including tenascin, thrombospondin, heparin, collagen and fibrinogen [50]. However its effects do not only rely on the molecular network with which fibronectin is connected. In fact, through the RGD motif, fibronectin binds to integrin receptors and in particular to integrin α5β1, which is markedly up-regulated during tumor-associated angiogenesis [51]. The deposition of fibronectin in a 3D cell-derived ECM appears to be imperative for matrix assembly and vascular morphogenesis [52]. Fibronectin prompts EC survival, whereas the blockage of its polymerization induces an impairment of EC proliferation and tube formation both in vitro and in vivo [48,53]. The importance of this molecule in angiogenesis is highlighted by the fact that, whereas the lack of the splice variants EIIIA or EIIIB does not lead to obvious defects, the double knockout mice show hemorrhages and severe vascular defects [54,55,56].
Given its wide distribution and interaction with other molecules involved in angiogenesis, it is likely that the altered expression of fibronectin observed in a number of pathologies such as fibrosis and cancer, may significantly impact on the development of new vessels and on the efficacy of anti-angiogenic therapy.

4. Collagens: A Major Source of Anti-Angiogenic Fragments

The synthesis and deposition of various collagens is known to affect EC survival and vessel formation. Those principally involved in angiogenesis are: type I, type IV, type XV and type XVIII collagen.
Type I collagen is the main ECM constituent to which proliferating ECs are exposed in an injured tissue. The interaction of collagen I with α1β1, α2β1, ανβ3 and ανβ5 integrins induces the activation of Mitogen-Activated Protein (MAP) kinase pathway supporting EC survival [57,58]. Moreover, the binding of collagen I to α1β1 and α2β1 integrins, suppresses c Adenosine Monophosphate (cAMP)-dependent Protein Kinase A (PKA), resulting in the reorganization of the actin fibers and in cell shape changes [59]. Collagen I is also important for lumen formation since it is required for the coalescence of pinocytic intracellular vacuoles [60,61]. Thus, a major role of this collagen can be recognized in the first stages of the angiogenic process.
Type IV collagen is one of the major constituents of the basement membrane [62,63], and is composed of six genetically distinct α chains (α 1–6) [64]. EC adhesion and migration are prompted by the triple-helical part of collagen IV, whereas the non collagenous domain NC1 does not support EC migration. Instead, the NC1 domains compete with intact collagen IV for the binding to integrins, inhibiting EC proliferation and tube formation [65,66]. In the early stages of tumor development, collagen IV exerts a pro-angiogenic function, once cryptic domains are exposed by MMP cleavage [67]. In contrast, in the late stages, it displays anti-angiogenic properties due to the release of arresten, canstatin and tumstatin derived from the α1, α2 and α3 chains, respectively [68,69,70]. Arresten inhibits EC migration and tumor growth through the engagement of integrin α1β1 and inhibition of MAP kinase [65,71]. Similarly, canstatin also inhibits EC migration and tube formation [72] and impairs angiopoietin-1-induced angiogenesis and lymphangiogenesis [73]. The mechanisms of action involve the binding to integrins ανβ3 and ανβ5, and inhibition of Protein Kinase B (PKB) also known as Akt focal adhesion kinase (FAK) and mammalian Target of Rapamycin (mTOR) [74]. In addition, canstatin induces Fas-dependent EC death, similarly to what observed with TSPs, despite through different mechanisms [75]. Finally, tumstatin binds to ανβ3 integrin though two distinct sites and displays anti-angiogenic and anti-tumor activity [76,77]. Thus, as observed for other ECM molecules, collagen IV displays an opposite function once the angiogenic stimulus is turned on and the proteases are activated. This likely represents a feedback mechanism able to halt the persistence of the pro-angiogenic signals. It can also be pointed out that the various fragments activate different molecular pathways thus amplifying the mechanisms of action.
Type XV collagen is a component of the vascular basement membrane [78]. Its carboxy-terminal globular Non-Collagenous (NC) domain called restin displays a high homology with endostatin and inhibits EC migration [79,80,81]. Type XV collagen deficient mice develop normally but display collapsed capillaries and EC degeneration in the heart and skeletal muscle, and this suggests that the molecule could exert an important role in microvessels stabilization [82].
Type XVIII collagen is found in different epithelial and vascular basement membranes. The proteolitic cleavage of the NC1 domain of the α1 chain by cathepsin L, B and K or MMPs releases the anti-angiogenic fragment endostatin [83,84]. Endostatin significantly reduces EC invasion in vitro through the blockage of matrix metalloproteinase 2 (MMP-2) [85]. The fragment also inhibits FGF-2 and VEGF-induced EC proliferation and migration [86,87]. Relying on its anti-angiogenic properties, endostatin displays potent anti-tumoral effects [88,89,90,91]. Remarkably, the treatment with the recombinant fragment induces vessel normalization and, as a consequence an improved chemotherapy efficacy [92]. Clinical trials have also been attempted with seesawing results: some deluded the expectations [93,94], whereas others demonstrated a potential benefit [91,95]. It is possible that this may depend on the tumor type and the quality of the vasculature characterizing each tumor. In addition, the microenvironment of the specific tumors and its intrinsic matrix composition may also impinge of the activity of endostatin.
Taken together, first the importance of this family of molecules in angiogenesis is highlighted by the fact that some of the members are intrinsic components of the vascular basement membrane. Given that vessel integrity affects vascular efficacy, an altered expression of these molecules may profoundly affect drug delivery and efficacy. Second, the pivotal role of this family in this context is that it represents an important source of fragments displaying strong angiogenic activities. As discussed for TSPs, these fragments represent a vital reservoir for the development of new drugs that may display high anti-angiogenic activity and low toxicity.

5. Laminins: Multiple Chains for Multiple Functions

The expression of laminins varies between cell types and during development [96]. Several sequences within the α, β and γ chains of laminins play a role in angiogenesis. The activation of MMP-9 following hypoxia leads to the degradation of laminin, which associates with an enhanced vascular pruning [97]. The RGD and IKVAV peptides present within the laminin sequence display important angiogenic functions, but to exert their activity they need to be unmasked by proteases [98]. The α4 (in particular laminin 411, α4:β1:γ1 and laminin 421, α4:β1:γ1), and α5 chains of laminin are the most abundant in the vascular basement membranes and thus play a major role in this context [99,100].
In particular, the α4 chain of laminin represents a high affinity ligand for ανβ3, α3β1 and α6β1 integrins and is required for EC adhesion and migration [101]. Interestingly, the α4 chain and its receptor Melanoma Cell Adhesion Molecule (MCAM) are highly expressed in the blood vessels of renal cell carcinomas and the levels of expression may serve to predict the patients’ outcome [102]. The importance of laminin 411 is also highlighted by the fact that, by activating the Notch-δ-like 4 signaling pathway, may regulate tip cell formation and affect vascular density [103]. On the other hand, laminin α5 chain-based synthetic peptides inhibit angiogenesis and block cell binding to FGF-2 [104]. Other studies report a complementary function for laminin α4 and α5: laminin α4 knockout mice show bleeding during embryogenesis resulting in a perinatal anemia, which is rescued 3 to 4 weeks postnatally once laminin α5 expression is turned on [105]. It has also been suggested that laminin 511 is required to maintain the β1 integrin-dependent anchorage of ECs in shear stress conditions and thus may be important to the formation of functional vessels [106,107]. In 3D EC cultures laminin is required for EC aggregation in end-to-end networks and the effect is α6 integrin dependent; the engagement of α6 integrin by laminin increases Vascular Endothelial Growth Factor Receptor 2 (VEGFR2) expression and Vascular Endothelial Growth Factor (VEGF) uptake by EC [108].
Laminin exerts an important role also during EC sprouting, in fact it has been recently demonstrated that, once bound to integrin α6β1, the receptor is sequestered from the podosome rosettes, key structures in sprouting angiogenesis [109].
Also for this family of molecules, characterized by different chain arrangements, the effect on angiogenesis is complex and occurs through the interaction and/or activation of major signaling pathways in ECs. As suggested by the results obtained with the use of the knockout models, the different molecules may display overlapping functions. Interestingly, since these molecules affect EC tip formation and sprouting angiogenesis they may be required in the early stages of angiogenesis. On the other side, the presence of these molecules in the blood vessels’ basement membrane suggests a role in the maintenance of vascular homeostasis.

6. Proteoglycans: Complex Functions from the Protein Core and Carbohydrate Chains

Proteoglycans (PGs) are complex macromolecules composed of a core protein carrying one or more covalently linked glycosaminoglycan (GAG) chains, whose number and type may differ to a great extent [110,111]. PGs perform multiple functions in cancer and angiogenesis by virtue of their polyhedric nature and their ability to interact with different ligands and receptors [112,113,114]. During tumor development and growth, PG expression is markedly modified within the tumor microenvironment impacting on cancer cell signaling and angiogenesis [115]. Heparan sulfate PG (HSPG) such as syndecans and glypicans are regulators of cancer progression and angiogenesis and serve as biomarkers for early detection and/or as pharmacological targets [116,117,118]. Notably, high concentrations of condroitin sulfate PG (CSPG), in particular versican and decorin, have been reported in the tumor stroma [111,119,120]. The major PGs affecting tumor angiogenesis are described below.
Perlecan is a ubiquitous multimodular proteoglycan consisting of five domains sharing homology with growth factors, immunoglobulins and adhesion molecules [121,122,123]. Given its widespread distribution [124,125] and its ability to interact with various molecules perlecan regulates various biological processes. Either via the protein core or the heparan sulfate chains, perlecan binds other ECM components as laminin, fibronectin and ECM1 and several growth factors including Fibroblast Growth Factor 1 (FGF-1), FGF-2, FGF-7, FGF-9 and FGF-18, FGF-binding protein, Platelet-Derived Growth Factor (PDGF), Hepatocyte Growth Factor (HGF), activin A, VEGF and progranulin [126,127,128,129,130,131,132]. The central role of perlecan in angiogenesis has been demonstrated by several independent studies and in various in vitro and in vivo models [133,134,135]. Collectively, these studies indicate that perlecan exerts a pro-angiogenic function by binding and presenting VEGF-A, PDGF and various FGFs to their cognate receptors and modulating their activity [136,137,138,139,140,141,142]. During tumor angiogenesis and cancer growth the perlecan protein core is degraded by proteases and the heparan sulfate chains are removed by heparanases. As a consequence, a plethora of growth factors trapped within the molecule are released affecting angiogenesis and cancer progression [114,143]. Interestingly, perlecan expression is often deregulated during cancer progression, generally leading to enhanced tumor invasiveness [144,145,146,147]. This may likely depend both on its structural role in the vessels’ basements membranes and its direct role in affecting angiogenesis.
Endorepellin is the C-terminal processed form of perlecan and, in contrast to the pro-angiogenic N-terminal domain, displays an opposite function: it inhibits EC migration, capillary morphogenesis, and in vivo angiogenesis [148,149,150,151]. The angiostatic function of endorepellin is in part due to its interaction with α2β1 integrin and activation of a signaling cascade leading to dissolution of the actin cytoskeleton, disruption of focal adhesions and the inhibition of EC migration [149]. Furthermore, the engagement of α2β1 integrin by endorepellin triggers the activation of the tyrosine phosphatase Src Homology region 2 domain-containing Phosphatase-1 (SHP-1) which, in turn, dephosphorylates and inactivates various Receptor Tyrosine Kinases (RTKs) including VEGFR2 [152]. Endorepellin is also able to directly interact with VEGFR2 at the EC surface inducing the transcriptional repression of HIF-1α and VEGF-A [153]. The binding of endorepellin with VEGFR2 is concurrent with its binding to α2β1 integrin and the dual receptor antagonism concur in the inhibition of EC migration and blood vessel maturation [154,155]. More recently, it was also demonstrated that endorepellin evokes autophagy via VEGFR2 signaling contributing to the angiostatic function of the fragment [156,157]. Thus, similarly to what observed with collagens, the cleavage of perlecan gives rise to a fragment displaying an opposite function, which may serve to counterbalance the angiogenic stimulus.
Decorin is a member of the small leucine-rich proteoglycan (SLRP) family and its structural features enable the binding to numerous other ECM molecules, growth factors and cytokines modulating angiogenesis with different mechanisms [158]. The binding to type I, type VI collagen and fibronectin affects collagen fibril formation and modulates the ECM rigidity and stiffness, structural properties known to influence angiogenesis [159,160,161]. As other ECM molecule, decorin modulates the activity of pro-angiogenic growth factors including VEGF, PDGF, FGF, Insulin Growth Factor (IGF) and angiopoietin by sequestering them and influencing their availability [162,163,164,165]. Decorin controls also the bioavailability of TGF-β whose release depends on the activity of different proteases involved in angiogenesis [166,167]. Furthermore, soluble decorin is a high affinity antagonistic ligand for several key receptors tyrosine kinases including Epithelial Growth Factor Receptor (EGFR), HGF Receptor, IGF-1 Receptor and VEGFR2 [168]. The engagement of decorin with cell surface receptors can either activate or inhibit the function of the receptor, depending on the physiological state of the tissue [169]. More recently, decorin has emerged as a soluble pro-autophagic cue [170]. This effect is mediated via a direct interaction with VEGFR2 which causes activation of AMP kinase signaling that ultimately culminate in a Paternally-expressed gene 3 (Peg3)/Beclin1/microtubule-associated proteins 1 Light Chain 3A (LC3) dependent autophagic program [171,172]. Thus, depending on the microenvironment in which angiogenesis occurs, decorin can exhibit either a pro- or an anti-angiogenic activity [165]. Nevertheless, in tumor-associated angiogenesis and in various inflammatory processes, the antiangiogenic activity is predominant, providing a potential basis for the development of decorin-based therapies [173,174,175].
Biglycan shares about 65% overall homology with decorin. Through the binding to VEGFA and activation of VEGFR2 biglycan exerts a pro-angiogenic stimulus in fracture healing processes [176]. The pro-angiogenic function is also the result of its interaction with and inactivation of endostatin [177]. In colon cancer, the over-expression of biglycan induces VEGF up-regulation and angiogenesis and thus the molecule may be a promising target for the development of anti-angiogenic strategies [178]. Thus, the efficacy of anti-angiogenic therapy may depend also on the expression of biglycan in the tumor microenvironment.
The syndecan family comprises four distinct genes encoding single-pass transmembrane protein cores. The ectodomain of syndecans is natively disordered and this characteristic allows syndecans to interact with a variety of proteins and ligands, thereby providing enrichment in their biological function [114,179]. Syndecan-1 is a key regulator of angiogenesis and this mainly occurs through the interaction with IGF-1 Receptor. This interaction is essential for the crucial cross-talk taking place between VEGFR2, the ανβ3 integrin and VE-cadherin during angiogenesis [180]. Syndecan-1 also modulates the VEGF-VEGFR2 signaling thus promoting EC proliferation and survival [181,182]. In breast cancer, stromal Syndecan-1 promotes FGF-2 and VEGF signaling and its expression in patients correlates with both vessel density and total vessel area [183]. In contrast, syndecan-2 impairs angiogenesis in human microvascular ECs [184]. Syndecans’ levels at the cell surface are regulated by proteolytic cleavage. In myeloma cells it was shown that syndecan-1 is shed and the binding of VEGF through its heparan sulfate chains stimulated tumor angiogenesis further supporting cancer growth [185]. Similar to what observed with syndecan-1, the shed ectodomains of syndecan-2 and syndecan-4 modulate angiogenesis. Shed syndecan-2 inhibits angiogenesis via a paracrine interaction with Cluster of Differentiation 148 (CD148), which in turn deactivates the α1β1 and α2β1 integrins, two main angiogenesis receptors [186]. The ectodomain of syndecan-4 is cleaved by the secreted metalloprotease A Disintegrin And Metalloproinase with Thrombospondin motifs 1 (ADAMTS1). The ectodomain in turn induces an altered distribution of actin stress fibers and a significant decrease of Ras Homolog gene family member A (RhoA) activity, thus promoting EC migration and favoring angiogenesis [187,188].
Glypicans are cell-surface PGs that share similarity with syndecans and interact with a multitude of ECM proteins, chemokines, growth factors and receptors [112]. Glypican-1 is the most prevalent member of the family expressed in ECs and in the vascular system. It is frequently overexpressed in different human malignancies including pancreatic carcinoma and breast cancer [189]. Glypican-1 is highly expressed in gliomas and this may contribute to the malignancy of this highly angiogenic tumor [190]. Glypican acts as a co-receptor or promoter of many angiogenic growth factors, including VEGF, FGFs, PDGF, heparin-binding EGF (HB-EGF), HGF, and IGF-1 [191]. A recent study indicates that the delivery of glypican-1 by nanoliposomes to the site of ischemic injury could function as an enhancer for growth factor activity, thus improving the response to local angiogenic therapies for the treatment of ischemia [192]. This represents another important example of the potential employment of ECM molecules/fragments for therapy purposes.
Lumican is localized in the peripheral blood vessels of adult human lungs and in the thickened intima of the coronary artery and displays binding affinity for αν integrins [193,194]. Lumican expression by ECs increases during the resolution phase of angiogenesis, when vessels return to a state of angiostasis [195]. The angiostatic function of this molecule is exerted through the inhibition of integrin α2β1 activity and reduction of MMP-14 expression in ECs [196]. In different tumor models lumican over-expression impaired tumor growth due to a reduced vascular density [197] likely associated with Fas-induced EC apoptosis [198].
As summarized in this section, proteoglycans are multifaceted molecules whose protein core is composed by a variety of modules. They are composed of intricate sugar chains which enable these glycoproteins to sequester and modulate the activity of different growth factors. Their GPI-anchored, transmembrane or extracellular nature enables them to affect different stages of angiogenesis during EC migration to form new vessels. Not only proteoglycans themselves offer the potential to develop new tools to halt angiogenesis, but also the enzymes that cleave these molecules, thus liberating their activity, can be viewed as a possible therapeutic target.

7. Hyaluronan: Not Only a Mere Glue

HA is a large polysaccharide composed of repeating N-acetylglucosamine and glucuronic acid disaccharide units. Under physiological conditions, HA can be composed of up to 25,000 disaccharide units, which become smaller and more dispersed in a pathological status [199]. Numerous studies have shown that HA signaling plays an important role in angiogenesis, mainly by influencing EC behavior. The biophysical functions of HA vary depending on its molecular size which is firmly regulated by the concerted activities of biosynthetic and degradation processes. For instance, low molecular weight HA (LMW-HA, generally < 200 kDa) was shown to prompt inflammation and angiogenesis [200] and stimulate EC proliferation, motility and tubule formation [201]. These oligosaccharides are able to interact with two critical EC receptors: CD44, playing a major role in EC adhesion and proliferation, and RHAMM (Receptor for HA-Mediated Motility), essential for EC invasion [202,203,204]. The role of LMW-HA in affecting angiogenesis during wound healing, diabetic mellitus and endometriosis, suggests the possibility to modulate HA as a putative therapeutic approach [205,206]. In contrast, high molecular weight HA (HMW-HA) was reported to be anti-angiogenic [207,208] and immunosuppressive [209]. Despite these evidences, their role in angiogenesis is controversial since LMW-HA was also shown to inhibit angiogenesis impairing the CXC ligand 12 (CXCL12)/CXC Receptor 4 (CXCR4) pathway, as opposed to what observed with HMW-HA [210]. In addition, HMW-HA seems to stimulate angiogenesis cooperating with versican and inducing FGF-2 expression [211]. Thus, the angiogenic properties of LMW-HA and HMW-HA may vary depending on the composition of the microenvironment. Nevertheless, a high HA expression was shown to correlate with increased angiogenesis and poor prognosis [212] and the assessment of this molecule may represent an important mean for the development of new prognostic tools [199].
Thus, given the controversial effect of this polysaccharide in angiogenesis, further preclinical analyses must be carried out to better clarify their role in blood vessel formation and verify their putative potential in clinics.

8. The EDEN Family: Two Members with Opposite Functions

The acronym EDEN (EMI Domain ENdowed) designates a family of ECM glycoproteins characterized by the presence of a cystein-rich EMI domain at the N-terminus [213,214,215,216].
This family of proteins can be clustered in distinct groups based on the molecular characteristics and the arrangements of the domains. The first group comprises the molecules containing the EMI domain at the N-terminus, the distinctive domain of the family, followed by a coiled-coil region and a gC1q domain at the C-terminus. This group includes MULTIMERIN1 [217], MULTIMERIN2 [218], EMILIN1 [219], and EMILIN2 [220]. The second group comprises only one molecule, EMILIN3 which lacks the gC1q domain and is required for notochord formation [221,222].
The third cluster includes two genes, Emu1 and Emu2 small collagenous molecules which share with the family only the presence of EMI domain [222].
While some of these molecules are characterized by a wide distribution in most connective tissues, like EMILIN1 and EMILIN2, the expression of other molecules such as that of MULTIMERIN2 (MMRN2) is more restricted. Despite their molecular affinity, two of them, i.e. MMRN2 and EMILIN2, exert an opposite function during angiogenesis.
MMRN2, also known as EndoGlyx-1, was identified during a screening for new antigenic markers of the vascular endothelium and was found to be expressed only at the level of the blood vessel endothelium [223]. In neoplastic tissues MMRN2 is deposited along tumor capillaries and in the “hot spots” of neoangiogenesis [223].
The role of this molecule was first described in a paper by Lorenzon et al. [224], where it was found that MMRN2 primarily impinged on EC migration. This depends on its ability to sequester VEGF-A and inhibit VEGFR-2 activation [224,225]. The binding capability has been pinpointed in a region towards the N-terminus of the molecule [225]. The VEGF-binding activity primarily occurs through the carbohydrate chains, since their removal significantly compromises, the binding [225]. In addition, the molecule seems also to affect the redistribution of VEGFR2 receptor at the EC surface and to be the reservoir of different members of the VEGF family of cytokines including VEGF-C, VEGF-D and Placental Growth Factor (PlGF) [225]. When over-expressed in the tumor microenvironment the molecule and its active fragment display a potent anti-tumoral effect [224,225]. These investigations suggest that the activity is indirect and hinges on an impaired vascularization. A positive role in affecting sprouting angiogenesis has also been reported, thanks to the interaction with the tumor endothelial marker CLEC14A [226,227]. This indicates that the molecule may display different effects during angiogenesis, depending on the stage of vessels development. Anyhow, given the presence of the molecule along all the blood vessels it is conceivable that it might exert a homeostatic role, thus halting the development of new vessels unless a strong pro-angiogenic signal is released. Under these conditions, the protein could be degraded by the proteases secreted by ECs during migration to allow an efficient sprouting. Preliminary results generated in our laboratory support this hypothesis, still we have only begun to scratch the surface to assess the role of this molecule in angiogenesis and vessel homeostasis (Figure 2).
EMILIN2 was cloned following a two-yeast hybrid screening and is characterized by a proline-rich segment between the coiled-coil region and the collagenous stalk [220]. EMILIN2 displays a low expression in adult aorta, small intestine and appendix, whereas the highest levels of the protein can be detected in fetal heart starting at E8.5 and reaching the highest levels at E11.5 and adult lung mice [228]. At the functional level, EMILIN2 significantly impairs the growth of a number of tumor types inducing tumor cell apoptotic death and impairing Wnt signaling [11,15]. The Emilin2 gene is frequently methylated in breast, lung and colorectal tumors and this suppression correlates with poor clinical outcome and increased limph node metastasis in breast cancer patients [229].
Interestingly, and unexpectedly, EMILIN-2 also stimulates the development of new vessels [230,231]. The molecular mechanisms by which EMILIN2 affect ECs behavior are being investigated in our laboratory and involve the over-production of cytokines which, in turn, promote EC proliferation and migration (Figure 3). Furthermore, tumor vessels developed in EMILIN2-deprived microenvironments display a worsen integrity of the basal lamina suggesting that it might also affect vessel perfusion and drug delivery (unpublished observations).
Thus, despite the similar domain arrangements, EMILIN2 and MMRN2 exert opposite functions. This may depend on the fact that the angiogenic activity of MMRN2 occurs through the coiled-coil region, which shares low homology with that of EMILIN2, of which the pro-angiogenic region has not been identified yet.

9. The CCN Family of Proteins as Regulators of Vascular Development and Pathological Angiogenesis

The Connective tissue growth factor Cystein rich protein and Nephroblastoma overexpressed gene (CCN) family of proteins includes six members (CCN1–CCN6) that share conserved functional domains. These proteins have been shown to be reservoirs of growth factors and to promote intracellular signaling. This occurs through the interaction with cell surface integrins, receptors, or other ECM molecules [232]. These molecules have been extensively studied and their role in the modulation of the proliferation, migration and adhesion of EC, among other cells, established. Despite the homology, the function of the different CCN proteins is exclusive due to the specific expression patterns [233]. For this reason, their correct deposition in many physiologic processes is essential and an unbalanced secretion of these molecules often leads to severe disorders contributing to cancer progression and the onset of vascular diseases.
The CCN1 (CYR61) protein is expressed by ECs and vascular smooth muscle cells (VMSC); cardiac expression occurs at E8.5 and persists until E11.5 during mouse embryo development. The importance of this molecule in angiogenesis is highlighted by the fact that CCN1 null mice die at E14.5 due to vascular defects [234]. The mechanism of action involves the engagement of integrin αvβ3 and the promotion of EC adhesion and migration [235]. CCN1 stimulates tumor growth and is associated with an increased intra-tumor vascularization [235]. By promoting the differentiation of progenitor ECs, CCN1 supports the re-endothelialization after vascular injury [236]. In addition, by targeting VEGF, Src homology 2 domain phosphatase-1 and Notch signaling, CCN1 affects the development of retinal vessels [237].
The CCN2 protein (CTGF) shares with CCN1 similar expression patterns and it is not only expressed by ECs and VSMC but also by pericytes and regulates the interaction of these cells with ECs [238]. CCN2 null mice die shortly after birth due to severe skeletal and vascular defects associated with an impaired pericyte recruitment and basement membrane organization [239]. Given the importance of pericytes in maintaining vascular stability and affecting their efficiency, it is likely that an altered expression of the molecule may also affect drug delivery. CCN2 induces a HIF-1α-dependent VEGF expression and this occurs through the up-regulation of miR-210 via the activation of Phosphoinositol 3 kinase (PI3K), PKB, extracellular signal-regulated kinase (ERK) and nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB)/ETS-domain containing protein 1 (ELK1), leading to increased angiogenesis [240].
The CCN3 (NOV) protein is structurally similar to CCN1 and 2 but displays a divergent function protecting from aberrant excessive vessel growth [241]. CCN3 supports EC adhesion and/or migration through integrins ανβ3, α5β1, α6β1 and through heparan sulfate proteoglycans inducing corneal vascularization [242]. It was recently demonstrated that CCN3 plays also a key role in the development of abdominal aortic aneurysm [243]. CCN3 stimulates the expression of VEGF in prostate cancer cells activating the focal adhesion kinase (FAK)/PKB/NF-κB signaling pathway. Furthermore, CCN3 stimulates the recruitment of macrophages which are skewed to an M2 phenotype, thus leading to enhanced tumor angiogenesis [244]. By modifying this key cellular microenvironmental component CCN3 is thus able to deeply affect cancer growth both directly, affecting cancer cell behavior, and indirectly, affecting angiogenesis.
The CCN4 (WISP1) protein is a major regulator of skeletal development and only recently a role in VSMC migration and proliferation has been recognized [245]. A stimulation of VEGF-A expression was also demonstrated for CCN4 and it occurs via the engagement of integrin αvβ3 and the consequent activation of the FAK/c-Src pathway. This leads to the transactivation of the EGFR/ERK/HIF1-α, which prompts the development of new blood vessels [246].
The CCN5 (WISP2) protein is expressed early and persists throughout embryonic development; it is expressed by EC, VSMC and heart myocardium suggesting a role in vascular function [247]. Indeed, recent evidences indicate that CCN5 exerts potent anti-angiogenic effects in an aortic ring vessel outgrowth model, and this angiostatic activity is abrogated by its cleavage [248]. Thus, its cleavage may be a mechanism through which, under a strong angiogenic stimulus, the potent anti-angiogenic properties of CCN5 are turned off.
Unlike all the other members of the family, CCN6 (WISP3) has not so far been demonstrated to play a role in vascular development.
Also for this family of molecules the evidences indicating an involvement in angiogenesis are growing. Given their dual action on cancer cells and vascular cells they thus represent a promising source for the development of new anti-tumor drugs. As indicated above CCN2, CCN3 and CCN4 induce VEGF activation. Thus, in case of high expression levels of these molecules, it should be taken into account that anti-angiogenic therapies aimed at blocking VEGF activity may be hampered.

10. Conclusions

In this review, we report same examples of the intricate and complex regulations exerted by components of the ECM and deeply affecting the angiogenic process. For this reason, one can conceive that the concentration and processing of these molecules in the tumor microenvironment could be taken into account as an approach to predict the therapeutic efficacy. In fact, through their action these molecules can impinge on drug delivery and efficacy by modifying the vessels’ efficiency. In addition, for their prominent role in vessel formation the ECM composition should be taken into account to predict the efficacy of anti-angiogenic approaches. Finally, their endogenous nature offers the potential for the development of new more efficacious and less toxic treatments of angiogenesis-dependent diseases.

Acknowledgments

We gratefully thank the Associazione Italiana per la Ricerca sul Cancro (AIRC, grant # IG-2012-12718 to Maurizio Mongiat) and the Italian Ministry of Health (grant # RF-2010-2312580 to Maurizio Mongiat) for their support.

Author Contributions

Maurizio Mongiat, Eva Andreuzzi and Alice Paulitti wrote and edited the manuscript, Giulia Tarticchio created the figures.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. List reporting the main ECM molecules involved in angiogenesis.
Table A1. List reporting the main ECM molecules involved in angiogenesis.
ECM ProteinsFragmentsAnti-Angiogenic ActivityPro-Angiogenic ActivityReferences
Trombospondins
TSP-1 [27,28,30,32,33,34,35,36,37,38,39,40,41,42,43,44]
TSP-2 [29,31,44]
Fibronectin
Fibronectin [51,52,53,54,55,56]
Collagens
Type I [58,59,60,61]
Type IV [67]
arresten [65,71]
canstatin [72,73,74,75]
tumstatin [76,77]
Type XVrestin [80,81,82]
Type XVIIIendostatin [81,83,84,85,86,87,92]
Laminins
Laminin 411 and 421 [101,102,103,105]
Laminin 511 [104,106]
Proteoglicans
Perlecan [112,133,134,135,136,137,138,139,140,141,142]
endorepellin [148,149,150,151,152,153,154,155,156,157]
Decorin [162,163,164,165,166,167,168,169,170,171,172,173,174,175]
[159,160,161,163,164,165,169,170,171,172,173,174,175]
Biglycan [176,177,178]
Syndecan 1 [181,182,183,185]
Syndecan 2 [184,186]
Syndecan 4 [187,188]
Glypicans [189,190,191,192]
Lumican [196,197,198]
Hyaluronan
LMW-HA [200,201,202,203,204,205,206]
[210]
HMW-HA [207,208]
[211]
EDEN Family
Multimerin 2 [224,225]
[226,227]
Δ2 fragment [225]
Emilin 2 [230,231]
CCN Family
CCN1 [235,236,237]
CCN2 [239,240]
CCN3 [241,242,243,244]
CCN4 [245,246]
CCN5 [248]
ECM: Extracellular matrix; TSP: Thrombospondin; LMW-HA: Low Molecular Weight Hyaluronan; HMW-HA: High Molecular Weight Hyaluronan; EDEN: EMI Domain ENdowed; CCN: Connective tissue growth factor Cystein rich protein and Nephroblastoma overexpressed gene.

References

  1. Ozbek, S.; Balasubramanian, P.G.; Chiquet-Ehrismann, R.; Tucker, R.P.; Adams, J.C. The evolution of extracellular matrix. Mol. Biol. Cell 2010, 21, 4300–4305. [Google Scholar] [CrossRef] [PubMed]
  2. Hynes, R.O. The extracellular matrix: Not just pretty fibrils. Science 2009, 326, 1216–1219. [Google Scholar] [CrossRef] [PubMed]
  3. Lu, P.; Takai, K.; Weaver, V.M.; Werb, Z. Extracellular matrix degradation and remodeling in development and disease. Cold Spring Harb. Perspect. Biol. 2011, 3. [Google Scholar] [CrossRef] [PubMed]
  4. Page-McCaw, A.; Ewald, A.J.; Werb, Z. Matrix metalloproteinases and the regulation of tissue remodelling. Nat. Rev. Mol. Cell Biol. 2007, 8, 221–233. [Google Scholar] [CrossRef] [PubMed]
  5. Cox, T.R.; Erler, J.T. Remodeling and homeostasis of the extracellular matrix: Implications for fibrotic diseases and cancer. Dis. Model. Mech. 2011, 4, 165–178. [Google Scholar] [CrossRef] [PubMed]
  6. Marastoni, S.; Ligresti, G.; Lorenzon, E.; Colombatti, A.; Mongiat, M. Extracellular matrix: A matter of life and death. Connect. Tissue Res. 2008, 49, 203–206. [Google Scholar] [CrossRef] [PubMed]
  7. Lu, P.; Weaver, V.M.; Werb, Z. The extracellular matrix: A dynamic niche in cancer progression. J. Cell Biol. 2012, 196, 395–406. [Google Scholar] [CrossRef] [PubMed]
  8. Sherman-Baust, C.A.; Weeraratna, A.T.; Rangel, L.B.; Pizer, E.S.; Cho, K.R.; Schwartz, D.R.; Shock, T.; Morin, P.J. Remodeling of the extracellular matrix through overexpression of collagen VI contributes to cisplatin resistance in ovarian cancer cells. Cancer Cell 2003, 3, 377–386. [Google Scholar] [CrossRef]
  9. Paszek, M.J.; Zahir, N.; Johnson, K.R.; Lakins, J.N.; Rozenberg, G.I.; Gefen, A.; Reinhart-King, C.A.; Margulies, S.S.; Dembo, M.; Boettiger, D.; et al. Tensional homeostasis and the malignant phenotype. Cancer Cell 2005, 8, 241–254. [Google Scholar] [CrossRef] [PubMed]
  10. DuFort, C.C.; Paszek, M.J.; Weaver, V.M. Balancing forces: Architectural control of mechanotransduction. Nat. Rev. Mol. Cell Biol. 2011, 12, 308–319. [Google Scholar] [CrossRef] [PubMed]
  11. Marastoni, S.; Andreuzzi, E.; Paulitti, A.; Colladel, R.; Pellicani, R.; Todaro, F.; Schiavinato, A.; Bonaldo, P.; Colombatti, A.; Mongiat, M. EMILIN2 down-modulates the Wnt signalling pathway and suppresses breast cancer cell growth and migration. J. Pathol. 2014, 232, 391–404. [Google Scholar] [CrossRef] [PubMed]
  12. Tai, I.T.; Tang, M.J. SPARC in cancer biology: Its role in cancer progression and potential for therapy. Drug Resist. Updates 2008, 11, 231–246. [Google Scholar] [CrossRef] [PubMed]
  13. Dhar, A.; Ray, A. The CCN family proteins in carcinogenesis. Exp. Oncol. 2010, 32, 2–9. [Google Scholar] [PubMed]
  14. Reed, C.C.; Waterhouse, A.; Kirby, S.; Kay, P.; Owens, R.T.; McQuillan, D.J.; Iozzo, R.V. Decorin prevents metastatic spreading of breast cancer. Oncogene 2005, 24, 1104–1110. [Google Scholar] [CrossRef] [PubMed]
  15. Mongiat, M.; Ligresti, G.; Marastoni, S.; Lorenzon, E.; Doliana, R.; Colombatti, A. Regulation of the extrinsic apoptotic pathway by the extracellular matrix glycoprotein EMILIN2. Mol. Cell. Biol. 2007, 27, 7176–7187. [Google Scholar] [CrossRef] [PubMed]
  16. Orend, G.; Chiquet-Ehrismann, R. Tenascin-C induced signaling in cancer. Cancer Lett. 2006, 244, 143–163. [Google Scholar] [CrossRef] [PubMed]
  17. Folkman, J. Angiogenesis. Annu. Rev. Med. 2006, 57, 1–18. [Google Scholar] [CrossRef] [PubMed]
  18. Carmeliet, P. Angiogenesis in health and disease. Nat. Med. 2003, 9, 653–660. [Google Scholar] [CrossRef] [PubMed]
  19. Jain, R.K. Normalization of tumor vasculature: An emerging concept in antiangiogenic therapy. Science 2005, 307, 58–62. [Google Scholar] [CrossRef] [PubMed]
  20. Wong, P.P.; Demircioglu, F.; Ghazaly, E.; Alrawashdeh, W.; Stratford, M.R.; Scudamore, C.L.; Cereser, B.; Crnogorac-Jurcevic, T.; McDonald, S.; Elia, G.; et al. Dual-Action Combination Therapy Enhances Angiogenesis while Reducing Tumor Growth and Spread. Cancer Cell 2015, 27, 123–137. [Google Scholar] [CrossRef] [PubMed]
  21. Andreoli, C.M.; Miller, J.W. Anti-vascular endothelial growth factor therapy for ocular neovascular disease. Curr. Opin. Ophthalmol. 2007, 18, 502–508. [Google Scholar] [CrossRef] [PubMed]
  22. Ausprunk, D.H.; Folkman, J. Migration and proliferation of endothelial cells in preformed and newly formed blood vessels during tumor angiogenesis. Microvasc. Res. 1977, 14, 53–65. [Google Scholar] [CrossRef]
  23. Dejana, E.; Languino, L.R.; Polentarutti, N.; Balconi, G.; Ryckewaert, J.J.; Larrieu, M.J.; Donati, M.B.; Mantovani, A.; Marguerie, G. Interaction between fibrinogen and cultured endothelial cells. Induction of migration and specific binding. J. Clin. Investig. 1985, 75, 11–18. [Google Scholar] [CrossRef] [PubMed]
  24. Senger, D.R.; Perruzzi, C.A.; Streit, M.; Koteliansky, V.E.; de Fougerolles, A.R.; Detmar, M. The α1β1 and α2β1 integrins provide critical support for vascular endothelial growth factor signaling, endothelial cell migration, and tumor angiogenesis. Am. J. Pathol. 2002, 160, 195–204. [Google Scholar] [CrossRef]
  25. Ricard-Blum, S.; Salza, R. Matricryptins and matrikines: Biologically active fragments of the extracellular matrix. Exp. Dermatol. 2014, 23, 457–463. [Google Scholar] [CrossRef] [PubMed]
  26. Adams, J.C.; Lawler, J. The thrombospondins. Cold Spring Harb. Perspect. Biol. 2011, 3, a009712. [Google Scholar] [CrossRef] [PubMed]
  27. Good, D.J.; Polverini, P.J.; Rastinejad, F.; Le Beau, M.M.; Lemons, R.S.; Frazier, W.A.; Bouck, N.P. A tumor suppressor-dependent inhibitor of angiogenesis is immunologically and functionally indistinguishable from a fragment of thrombospondin. Proc. Natl. Acad. Sci. USA 1990, 87, 6624–6628. [Google Scholar] [CrossRef] [PubMed]
  28. Tolsma, S.S.; Volpert, O.V.; Good, D.J.; Frazier, W.A.; Polverini, P.J.; Bouck, N. Peptides derived from two separate domains of the matrix protein thrombospondin-1 have anti-angiogenic activity. J. Cell Biol. 1993, 122, 497–511. [Google Scholar] [CrossRef] [PubMed]
  29. Kyriakides, T.R.; Zhu, Y.H.; Smith, L.T.; Bain, S.D.; Yang, Z.; Lin, M.T.; Danielson, K.G.; Iozzo, R.V.; LaMarca, M.; McKinney, C.E.; et al. Mice that lack thrombospondin 2 display connective tissue abnormalities that are associated with disordered collagen fibrillogenesis, an increased vascular density, and a bleeding diathesis. J. Cell Biol. 1998, 140, 419–430. [Google Scholar] [CrossRef] [PubMed]
  30. Bleuel, K.; Popp, S.; Fusenig, N.E.; Stanbridge, E.J.; Boukamp, P. Tumor suppression in human skin carcinoma cells by chromosome 15 transfer or thrombospondin-1 overexpression through halted tumor vascularization. Proc. Natl. Acad. Sci. USA 1999, 96, 2065–2070. [Google Scholar] [CrossRef] [PubMed]
  31. Streit, M.; Riccardi, L.; Velasco, P.; Brown, L.F.; Hawighorst, T.; Bornstein, P.; Detmar, M. Thrombospondin-2: A potent endogenous inhibitor of tumor growth and angiogenesis. Proc. Natl. Acad. Sci. USA 1999, 96, 14888–14893. [Google Scholar] [CrossRef] [PubMed]
  32. Streit, M.; Velasco, P.; Brown, L.F.; Skobe, M.; Richard, L.; Riccardi, L.; Lawler, J.; Detmar, M. Overexpression of thrombospondin-1 decreases angiogenesis and inhibits the growth of human cutaneous squamous cell carcinomas. Am. J. Pathol. 1999, 155, 441–452. [Google Scholar] [CrossRef]
  33. Weinstat-Saslow, D.L.; Zabrenetzky, V.S.; VanHoutte, K.; Frazier, W.A.; Roberts, D.D.; Steeg, P.S. Transfection of thrombospondin 1 complementary DNA into a human breast carcinoma cell line reduces primary tumor growth, metastatic potential, and angiogenesis. Cancer Res. 1994, 54, 6504–6511. [Google Scholar] [PubMed]
  34. Dawson, D.W.; Pearce, S.F.; Zhong, R.; Silverstein, R.L.; Frazier, W.A.; Bouck, N.P. CD36 mediates the In vitro inhibitory effects of thrombospondin-1 on endothelial cells. J. Cell Biol. 1997, 138, 707–717. [Google Scholar] [CrossRef] [PubMed]
  35. Dawson, D.W.; Volpert, O.V.; Pearce, S.F.; Schneider, A.J.; Silverstein, R.L.; Henkin, J.; Bouck, N.P. Three distinct D-amino acid substitutions confer potent antiangiogenic activity on an inactive peptide derived from a thrombospondin-1 type 1 repeat. Mol. Pharmacol. 1999, 55, 332–338. [Google Scholar] [PubMed]
  36. Yee, K.O.; Connolly, C.M.; Duquette, M.; Kazerounian, S.; Washington, R.; Lawler, J. The effect of thrombospondin-1 on breast cancer metastasis. Breast Cancer Res. Treat. 2009, 114, 85–96. [Google Scholar] [CrossRef] [PubMed]
  37. Asch, A.S.; Silbiger, S.; Heimer, E.; Nachman, R.L. Thrombospondin sequence motif (CSVTCG) is responsible for CD36 binding. Biochem. Biophys. Res. Commun. 1992, 182, 1208–1217. [Google Scholar] [CrossRef]
  38. Jimenez, B.; Volpert, O.V.; Reiher, F.; Chang, L.; Munoz, A.; Karin, M.; Bouck, N. C-Jun N-terminal kinase activation is required for the inhibition of neovascularization by thrombospondin-1. Oncogene 2001, 20, 3443–3448. [Google Scholar] [CrossRef] [PubMed]
  39. Belotti, D.; Foglieni, C.; Resovi, A.; Giavazzi, R.; Taraboletti, G. Targeting angiogenesis with compounds from the extracellular matrix. Int. J. Biochem. Cell Biol. 2011, 43, 1674–1685. [Google Scholar] [CrossRef] [PubMed]
  40. Lawler, P.R.; Lawler, J. Molecular basis for the regulation of angiogenesis by thrombospondin-1 and -2. Cold Spring Harb. Perspect. Med. 2012, 2, a006627. [Google Scholar] [CrossRef] [PubMed]
  41. Resovi, A.; Pinessi, D.; Chiorino, G.; Taraboletti, G. Current understanding of the thrombospondin-1 interactome. Matrix Biol. 2014, 37, 83–91. [Google Scholar] [CrossRef] [PubMed]
  42. Taraboletti, G.; Rusnati, M.; Ragona, L.; Colombo, G. Targeting tumor angiogenesis with TSP-1-based compounds: Rational design of antiangiogenic mimetics of endogenous inhibitors. Oncotarget 2010, 1, 662–673. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, X.; Lawler, J. Thrombospondin-based antiangiogenic therapy. Microvasc. Res. 2007, 74, 90–99. [Google Scholar] [CrossRef] [PubMed]
  44. Lawler, J.; Detmar, M. Tumor progression: The effects of thrombospondin-1 and -2. Int. J. Biochem. Cell Biol. 2004, 36, 1038–1045. [Google Scholar] [CrossRef] [PubMed]
  45. Haviv, F.; Bradley, M.F.; Kalvin, D.M.; Schneider, A.J.; Davidson, D.J.; Majest, S.M.; McKay, L.M.; Haskell, C.J.; Bell, R.L.; Nguyen, B.; et al. Thrombospondin-1 Mimetic Peptide Inhibitors of Angiogenesis and Tumor Growth: Design, Synthesis, and Optimization of Pharmacokinetics and Biological Activities. J. Med. Chem. 2005, 48, 2838–2846. [Google Scholar] [CrossRef] [PubMed]
  46. Uronis, H.E.; Cushman, S.M.; Bendell, J.C.; Blobe, G.C.; Morse, M.A.; Nixon, A.B.; Dellinger, A.; Starr, M.D.; Li, H.; Meadows, K.; et al. A phase I study of ABT-510 plus bevacizumab in advanced solid tumors. Cancer Med. 2013, 2, 316–324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Peters, J.H.; Chen, G.E.; Hynes, R.O. Fibronectin isoform distribution in the mouse. II. Differential distribution of the alternatively spliced EIIIB, EIIIA, and V segments in the adult mouse. Cell Adhes. Commun. 1996, 4, 127–148. [Google Scholar] [CrossRef] [PubMed]
  48. Astrof, S.; Hynes, R.O. Fibronectins in vascular morphogenesis. Angiogenesis 2009, 12, 165–175. [Google Scholar] [CrossRef] [PubMed]
  49. Pedretti, M.; Soltermann, A.; Arni, S.; Weder, W.; Neri, D.; Hillinger, S. Comparative immunohistochemistry of L19 and F16 in non-small cell lung cancer and mesothelioma: Two human antibodies investigated in clinical trials in patients with cancer. Lung Cancer 2009, 64, 28–33. [Google Scholar] [CrossRef] [PubMed]
  50. Singh, P.; Carraher, C.; Schwarzbauer, J.E. Assembly of Fibronectin Extracellular Matrix. Annu. Rev. Cell Dev. Biol. 2010, 26, 397–419. [Google Scholar] [CrossRef] [PubMed]
  51. Hynes, R.O. Cell-matrix adhesion in vascular development. J. Thromb. Haemost. 2007, 5, 32–40. [Google Scholar] [CrossRef] [PubMed]
  52. Hielscher, A.; Ellis, K.; Qiu, C.; Porterfield, J.; Gerecht, S. Fibronectin Deposition Participates in Extracellular Matrix Assembly and Vascular Morphogenesis. PLoS ONE 2016, 11, e0147600. [Google Scholar] [CrossRef] [PubMed]
  53. Zhou, X.; Rowe, R.G.; Hiraoka, N.; George, J.P.; Wirtz, D.; Mosher, D.F.; Virtanen, I.; Chernousov, M.A.; Weiss, S.J. Fibronectin fibrillogenesis regulates three-dimensional neovessel formation. Genes Dev. 2008, 22, 1231–1243. [Google Scholar] [CrossRef] [PubMed]
  54. Tan, M.H.; Sun, Z.; Opitz, S.L.; Schmidt, T.E.; Peters, J.H.; George, E.L. Deletion of the alternatively spliced fibronectin EIIIA domain in mice reduces atherosclerosis. Blood 2004, 104, 11–18. [Google Scholar] [CrossRef] [PubMed]
  55. Fukuda, T.; Yoshida, N.; Kataoka, Y.; Manabe, R.; Mizuno-Horikawa, Y.; Sato, M.; Kuriyama, K.; Yasui, N.; Sekiguchi, K. Mice lacking the EDB segment of fibronectin develop normally but exhibit reduced cell growth and fibronectin matrix assembly in vitro. Cancer Res. 2002, 62, 5603–5610. [Google Scholar] [PubMed]
  56. Astrof, S.; Crowley, D.; Hynes, R.O. Multiple cardiovascular defects caused by the absence of alternatively spliced segments of fibronectin. Dev. Biol. 2007, 311, 11–24. [Google Scholar] [CrossRef] [PubMed]
  57. Perruzzi, C.A.; de Fougerolles, A.R.; Koteliansky, V.E.; Whelan, M.C.; Westlin, W.F.; Senger, D.R. Functional overlap and cooperativity among αv and β1 integrin subfamilies during skin angiogenesis. J. Investig. Dermatol. 2003, 120, 1100–1109. [Google Scholar] [CrossRef] [PubMed]
  58. Sweeney, S.M.; DiLullo, G.; Slater, S.J.; Martinez, J.; Iozzo, R.V.; Lauer-Fields, J.L.; Fields, G.B.; San Antonio, J.D. Angiogenesis in collagen I requires α2β1 ligation of a GFP*GER sequence and possibly p38 MAPK activation and focal adhesion disassembly. J. Biol. Chem. 2003, 278, 30516–30524. [Google Scholar] [CrossRef] [PubMed]
  59. Whelan, M.C.; Senger, D.R. Collagen I initiates endothelial cell morphogenesis by inducing actin polymerization through suppression of cyclic AMP and protein kinase A. J. Biol. Chem. 2003, 278, 327–334. [Google Scholar] [CrossRef] [PubMed]
  60. Kamei, M.; Saunders, W.B.; Bayless, K.J.; Dye, L.; Davis, G.E.; Weinstein, B.M. Endothelial tubes assemble from intracellular vacuoles in vivo. Nature 2006, 442, 453–456. [Google Scholar] [CrossRef] [PubMed]
  61. Stratman, A.N.; Saunders, W.B.; Sacharidou, A.; Koh, W.; Fisher, K.E.; Zawieja, D.C.; Davis, M.J.; Davis, G.E. Endothelial cell lumen and vascular guidance tunnel formation requires MT1-MMP-dependent proteolysis in 3-dimensional collagen matrices. Blood 2009, 114, 237–247. [Google Scholar] [CrossRef] [PubMed]
  62. Davis, G.E.; Senger, D.R. Endothelial extracellular matrix: Biosynthesis, remodeling, and functions during vascular morphogenesis and neovessel stabilization. Circ. Res. 2005, 97, 1093–1107. [Google Scholar] [CrossRef] [PubMed]
  63. De Smet, F.; Segura, I.; De, B.K.; Hohensinner, P.J.; Carmeliet, P. Mechanisms of vessel branching: Filopodia on endothelial tip cells lead the way. Arterioscler. Thromb. Vasc. Biol. 2009, 29, 639–649. [Google Scholar] [CrossRef] [PubMed]
  64. Mundel, T.M.; Kalluri, R. Type IV collagen-derived angiogenesis inhibitors. Microvasc. Res. 2007, 74, 85–89. [Google Scholar] [CrossRef] [PubMed]
  65. Sudhakar, A.; Nyberg, P.; Keshamouni, V.G.; Mannam, A.P.; Li, J.; Sugimoto, H.; Cosgrove, D.; Kalluri, R. Human α1 type IV collagen NC1 domain exhibits distinct antiangiogenic activity mediated by α1β1 integrin. J. Clin. Investig. 2005, 115, 2801–2810. [Google Scholar] [CrossRef] [PubMed]
  66. Kalluri, R. Discovery of type IV collagen non-collagenous domains as novel integrin ligands and endogenous inhibitors of angiogenesis. Cold Spring Harb. Symp. Quant. Biol. 2002, 67, 255–266. [Google Scholar] [CrossRef] [PubMed]
  67. Xu, J.; Rodriguez, D.; Petitclerc, E.; Kim, J.J.; Hangai, M.; Moon, Y.S.; Davis, G.E.; Brooks, P.C. Proteolytic exposure of a cryptic site within collagen type IV is required for angiogenesis and tumor growth in vivo. J. Cell Biol. 2001, 154, 1069–1079. [Google Scholar] [CrossRef] [PubMed]
  68. Kim, Y.M.; Jang, J.W.; Lee, O.H.; Yeon, J.; Choi, E.Y.; Kim, K.W.; Lee, S.T.; Kwon, Y.G. Endostatin inhibits endothelial and tumor cellular invasion by blocking the activation and catalytic activity of matrix metalloproteinase. Cancer Res. 2000, 60, 5410–5413. [Google Scholar] [PubMed]
  69. Kamphaus, G.D.; Colorado, P.C.; Panka, D.J.; Hopfer, H.; Ramchandran, R.; Torre, A.; Maeshima, Y.; Mier, J.W.; Sukhatme, V.P.; Kalluri, R. Canstatin, a Novel Matrix-derived Inhibitor of Angiogenesis and Tumor Growth. J. Biol. Chem. 2000, 275, 1209–1215. [Google Scholar] [CrossRef] [PubMed]
  70. Petitclerc, E.; Boutaud, A.; Prestayko, A.; Xu, J.; Sado, Y.; Ninomiya, Y.; Sarras, M.P., Jr.; Hudson, B.G.; Brooks, P.C. New functions for non-collagenous domains of human collagen type IV. Novel integrin ligands inhibiting angiogenesis and tumor growth in vivo. J. Biol. Chem. 2000, 275, 8051–8061. [Google Scholar] [CrossRef] [PubMed]
  71. Aikio, M.; Alahuhta, I.; Nurmenniemi, S.; Suojanen, J.; Palovuori, R.; Teppo, S.; Sorsa, T.; Lopez-Otin, C.; Pihlajaniemi, T.; Salo, T.; et al. Arresten, a collagen-derived angiogenesis inhibitor, suppresses invasion of squamous cell carcinoma. PLoS ONE 2012, 7, e51044. [Google Scholar] [CrossRef] [PubMed]
  72. He, G.A.; Luo, J.X.; Zhang, T.Y.; Wang, F.Y.; Li, R.F. Canstatin-N fragment inhibits in vitro endothelial cell proliferation and suppresses in vivo tumor growth. Biochem. Biophys. Res. Commun. 2003, 312, 801–805. [Google Scholar] [CrossRef] [PubMed]
  73. Hwang-Bo, J.; Yoo, K.H.; Park, J.H.; Jeong, H.S.; Chung, I.S. Recombinant canstatin inhibits angiopoietin-1-induced angiogenesis and lymphangiogenesis. Int. J. Cancer 2012, 131, 298–309. [Google Scholar] [CrossRef] [PubMed]
  74. Magnon, C.; Galaup, A.; Mullan, B.; Rouffiac, V.; Bouquet, C.; Bidart, J.M.; Griscelli, F.; Opolon, P.; Perricaudet, M. Canstatin acts on endothelial and tumor cells via mitochondrial damage initiated through interaction with αvβ3 and αvβ5 integrins. Cancer Res. 2005, 65, 4353–4361. [Google Scholar] [CrossRef] [PubMed]
  75. Panka, D.J.; Mier, J.W. Canstatin inhibits Akt activation and induces Fas-dependent apoptosis in endothelial cells. J. Biol. Chem. 2003, 278, 37632–37636. [Google Scholar] [CrossRef] [PubMed]
  76. Maeshima, Y.; Sudhakar, A.; Lively, J.C.; Ueki, K.; Kharbanda, S.; Kahn, C.R.; Sonenberg, N.; Hynes, R.O.; Kalluri, R. Tumstatin, an endothelial cell-specific inhibitor of protein synthesis. Science 2002, 295, 140–143. [Google Scholar] [CrossRef] [PubMed]
  77. Xie, L.; Duncan, M.B.; Pahler, J.; Sugimoto, H.; Martino, M.; Lively, J.; Mundel, T.; Soubasakos, M.; Rubin, K.; Takeda, T.; et al. Counterbalancing angiogenic regulatory factors control the rate of cancer progression and survival in a stage-specific manner. Proc. Natl. Acad. Sci. USA 2011, 108, 9939–9944. [Google Scholar] [CrossRef] [PubMed]
  78. Myers, J.C.; Dion, A.S.; Abraham, V.; Amenta, P.S. Type XV collagen exhibits a widespread distribution in human tissues but a distinct localization in basement membrane zones. Cell Tissue Res. 1996, 286, 493–505. [Google Scholar] [CrossRef] [PubMed]
  79. Muragaki, Y.; Abe, N.; Ninomiya, Y.; Olsen, B.R.; Ooshima, A. The human α 1(XV) collagen chain contains a large amino-terminal non-triple helical domain with a tandem repeat structure and homology to α 1(XVIII) collagen. J. Biol. Chem. 1994, 269, 4042–4046. [Google Scholar] [PubMed]
  80. Ramchandran, R.; Dhanabal, M.; Volk, R.; Waterman, M.J.; Segal, M.; Lu, H.; Knebelmann, B.; Sukhatme, V.P. Antiangiogenic activity of restin, NC10 domain of human collagen XV: Comparison to endostatin. Biochem. Biophys. Res. Commun. 1999, 255, 735–739. [Google Scholar] [CrossRef] [PubMed]
  81. John, H.; Radtke, K.; Standker, L.; Forssmann, W.G. Identification and characterization of novel endogenous proteolytic forms of the human angiogenesis inhibitors restin and endostatin. Biochim. Biophys. Acta 2005, 1747, 161–170. [Google Scholar] [CrossRef] [PubMed]
  82. Eklund, L.; Piuhola, J.; Komulainen, J.; Sormunen, R.; Ongvarrasopone, C.; Fassler, R.; Muona, A.; Ilves, M.; Ruskoaho, H.; Takala, T.E.; et al. Lack of type XV collagen causes a skeletal myopathy and cardiovascular defects in mice. Proc. Natl. Acad. Sci. USA 2001, 98, 1194–1199. [Google Scholar] [CrossRef] [PubMed]
  83. O’Reilly, M.S.; Boehm, T.; Shing, Y.; Fukai, N.; Vasios, G.; Lane, W.S.; Flynn, E.; Birkhead, J.R.; Olsen, B.R.; Folkman, J. Endostatin: An endogenous inhibitor of angiogenesis and tumor growth. Cell 1997, 88, 277–285. [Google Scholar] [CrossRef]
  84. Ferreras, M.; Felbor, U.; Lenhard, T.; Olsen, B.R.; Delaisse, J. Generation and degradation of human endostatin proteins by various proteinases. FEBS Lett. 2000, 486, 247–251. [Google Scholar] [CrossRef]
  85. Lee, S.J.; Jang, J.W.; Kim, Y.M.; Lee, H.I.; Jeon, J.Y.; Kwon, Y.G.; Lee, S.T. Endostatin binds to the catalytic domain of matrix metalloproteinase-2. FEBS Lett. 2002, 519, 147–152. [Google Scholar] [CrossRef]
  86. Chang, J.H.; Gabison, E.E.; Kato, T.; Azar, D.T. Corneal neovascularization. Curr. Opin. Ophthalmol. 2001, 12, 242–249. [Google Scholar] [CrossRef] [PubMed]
  87. Ling, Y.; Yang, Y.; Lu, N.; You, Q.D.; Wang, S.; Gao, Y.; Chen, Y.; Guo, Q.L. Endostar, a novel recombinant human endostatin, exerts antiangiogenic effect via blocking VEGF-induced tyrosine phosphorylation of KDR/Flk-1 of endothelial cells. Biochem. Biophys. Res. Commun. 2007, 361, 79–84. [Google Scholar] [CrossRef] [PubMed]
  88. Kim, H.S.; Lim, S.J.; Park, Y.K. Anti-angiogenic factor endostatin in osteosarcoma. APMIS 2009, 117, 716–723. [Google Scholar] [CrossRef] [PubMed]
  89. Chen, Y.; Huang, H.; Yao, C.; Su, F.; Guan, W.; Yan, S.; Ni, Z. Antitumor activity of combined endostatin and thymidine kinase gene therapy in C6 glioma models. Cancer Med. 2016, 5, 2477–2486. [Google Scholar] [CrossRef] [PubMed]
  90. Ferician, O.; Cimpean, A.M.; Avram, S.; Raica, M. Endostatin Effects on Tumor Cells and Vascular Network of Human Renal Cell Carcinoma Implanted on Chick Embryo Chorioallantoic Membrane. Anticancer Res. 2015, 35, 6521–6528. [Google Scholar] [PubMed]
  91. Guan, Y.; Li, A.; Xiao, W.; Liu, S.; Chen, B.; Lu, T.; Zhao, C.; Han, F. The efficacy and safety of Endostar combined with chemoradiotherapy for patients with advanced, locally recurrent nasopharyngeal carcinoma. Oncotarget 2015, 6, 33926–33934. [Google Scholar] [PubMed]
  92. Li, W.; Zhao, X.; Du, B.; Li, X.; Liu, S.; Yang, X.Y.; Ding, H.; Yang, W.; Pan, F.; Wu, X.; et al. Gold Nanoparticle-Mediated Targeted Delivery of Recombinant Human Endostatin Normalizes Tumour Vasculature and Improves Cancer Therapy. Sci. Rep. 2016, 6, 30619. [Google Scholar] [CrossRef] [PubMed]
  93. Sun, X.J.; Deng, Q.H.; Yu, X.M.; Ji, Y.L.; Zheng, Y.D.; Jiang, H.; Xu, Y.P.; Ma, S.L. A phase II study of Endostatin in combination with paclitaxel, carboplatin, and radiotherapy in patients with unresectable locally advanced non-small cell lung cancer. BMC Cancer 2016, 16, 266. [Google Scholar] [CrossRef] [PubMed]
  94. Kulke, M.H.; Bergsland, E.K.; Ryan, D.P.; Enzinger, P.C.; Lynch, T.J.; Zhu, A.X.; Meyerhardt, J.A.; Heymach, J.V.; Fogler, W.E.; Sidor, C.; et al. Phase II study of recombinant human endostatin in patients with advanced neuroendocrine tumors. J. Clin. Oncol. 2006, 24, 3555–3561. [Google Scholar] [CrossRef] [PubMed]
  95. Cui, C.; Mao, L.; Chi, Z.; Si, L.; Sheng, X.; Kong, Y.; Li, S.; Lian, B.; Gu, K.; Tao, M.; et al. A phase II, randomized, double-blind, placebo-controlled multicenter trial of Endostar in patients with metastatic melanoma. Mol. Ther. 2013, 21, 1456–1463. [Google Scholar] [CrossRef] [PubMed]
  96. Durbeej, M. Laminins. Cell Tissue Res. 2010, 339, 259–268. [Google Scholar] [CrossRef] [PubMed]
  97. Boroujerdi, A.; Welser-Alves, J.V.; Milner, R. Matrix metalloproteinase-9 mediates post-hypoxic vascular pruning of cerebral blood vessels by degrading laminin and claudin-5. Angiogenesis 2015, 18, 255–264. [Google Scholar] [CrossRef] [PubMed]
  98. Khan, K.M.; Falcone, D.J. Role of laminin in matrix induction of macrophage urokinase-type plasminogen activator and 92-kDa metalloproteinase expression. J. Biol. Chem. 1997, 272, 8270–8275. [Google Scholar] [CrossRef] [PubMed]
  99. Yousif, L.F.; di Russo, J.; Sorokin, L. Laminin isoforms in endothelial and perivascular basement membranes. Cell Adhes. Migr. 2013, 7, 101–110. [Google Scholar] [CrossRef] [PubMed]
  100. Sixt, M.; Engelhardt, B.; Pausch, F.; Hallmann, R.; Wendler, O.; Sorokin, L.M. Endothelial cell laminin isoforms, laminins 8 and 10, play decisive roles in T cell recruitment across the blood-brain barrier in experimental autoimmune encephalomyelitis. J. Cell Biol. 2001, 153, 933–946. [Google Scholar] [CrossRef] [PubMed]
  101. Gonzalez, A.M.; Gonzales, M.; Herron, G.S.; Nagavarapu, U.; Hopkinson, S.B.; Tsuruta, D.; Jones, J.C. Complex interactions between the laminin α4 subunit and integrins regulate endothelial cell behavior in vitro and angiogenesis in vivo. Proc. Natl. Acad. Sci. USA 2002, 99, 16075–16080. [Google Scholar] [CrossRef] [PubMed]
  102. Wragg, J.W.; Finnity, J.P.; Anderson, J.A.; Ferguson, H.J.; Porfiri, E.; Bhatt, R.I.; Murray, P.G.; Heath, V.L.; Bicknell, R. MCAM and LAMA4 Are Highly Enriched in Tumor Blood Vessels of Renal Cell Carcinoma and Predict Patient Outcome. Cancer Res. 2016, 76, 2314–2326. [Google Scholar] [CrossRef] [PubMed]
  103. Estrach, S.; Cailleteau, L.; Franco, C.A.; Gerhardt, H.; Stefani, C.; Lemichez, E.; Gagnoux-Palacios, L.; Meneguzzi, G.; Mettouchi, A. Laminin-binding integrins induce Dll4 expression and Notch signaling in endothelial cells. Circ. Res. 2011, 109, 172–182. [Google Scholar] [CrossRef] [PubMed]
  104. Hibino, S.; Shibuya, M.; Hoffman, M.P.; Engbring, J.A.; Hossain, R.; Mochizuki, M.; Kudoh, S.; Nomizu, M.; Kleinman, H.K. Laminin α5 chain metastasis- and angiogenesis-inhibiting peptide blocks fibroblast growth factor 2 activity by binding to the heparan sulfate chains of CD44. Cancer Res. 2005, 65, 10494–10501. [Google Scholar] [CrossRef] [PubMed]
  105. Thyboll, J.; Kortesmaa, J.; Cao, R.; Soininen, R.; Wang, L.; Iivanainen, A.; Sorokin, L.; Risling, M.; Cao, Y.; Tryggvason, K. Deletion of the laminin α4 chain leads to impaired microvessel maturation. Mol. Cell. Biol. 2002, 22, 1194–1202. [Google Scholar] [CrossRef] [PubMed]
  106. Di Russo, J.; Hannocks, M.J.; Luik, A.L.; Song, J.; Zhang, X.; Yousif, L.; Aspite, G.; Hallmann, R.; Sorokin, L. Vascular laminins in physiology and pathology. Matrix Biol. 2016. [Google Scholar] [CrossRef] [PubMed]
  107. Yamamoto, H.; Ehling, M.; Kato, K.; Kanai, K.; van Lessen, M.; Frye, M.; Zeuschner, D.; Nakayama, M.; Vestweber, D.; Adams, R.H. Integrin β1 controls VE-cadherin localization and blood vessel stability. Nat. Commun. 2015, 6, 6429. [Google Scholar] [CrossRef] [PubMed]
  108. Stamati, K.; Priestley, J.V.; Mudera, V.; Cheema, U. Laminin promotes vascular network formation in 3D in vitro collagen scaffolds by regulating VEGF uptake. Exp. Cell Res. 2014, 327, 68–77. [Google Scholar] [CrossRef] [PubMed]
  109. Seano, G.; Chiaverina, G.; Gagliardi, P.A.; di Blasio, L.; Puliafito, A.; Bouvard, C.; Sessa, R.; Tarone, G.; Sorokin, L.; Helley, D.; et al. Endothelial podosome rosettes regulate vascular branching in tumour angiogenesis. Nat. Cell Biol. 2014, 16, 931–938. [Google Scholar] [CrossRef] [PubMed]
  110. Theocharis, A.D.; Skandalis, S.S.; Tzanakakis, G.N.; Karamanos, N.K. Proteoglycans in health and disease: Novel roles for proteoglycans in malignancy and their pharmacological targeting. FEBS J. 2010, 277, 3904–3923. [Google Scholar] [CrossRef] [PubMed]
  111. Afratis, N.; Gialeli, C.; Nikitovic, D.; Tsegenidis, T.; Karousou, E.; Theocharis, A.D.; Pavao, M.S.; Tzanakakis, G.N.; Karamanos, N.K. Glycosaminoglycans: Key players in cancer cell biology and treatment. FEBS J. 2012, 279, 1177–1197. [Google Scholar] [CrossRef] [PubMed]
  112. Iozzo, R.V.; Sanderson, R.D. Proteoglycans in cancer biology, tumour microenvironment and angiogenesis. J. Cell. Mol. Med. 2011, 15, 1013–1031. [Google Scholar] [CrossRef] [PubMed]
  113. Chiodelli, P.; Bugatti, A.; Urbinati, C.; Rusnati, M. Heparin/Heparan sulfate proteoglycans glycomic interactome in angiogenesis: Biological implications and therapeutical use. Molecules 2015, 20, 6342–6388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Iozzo, R.V.; Schaefer, L. Proteoglycan form and function: A comprehensive nomenclature of proteoglycans. Matrix Biol. 2015, 42, 11–55. [Google Scholar] [CrossRef] [PubMed]
  115. Theocharis, A.D.; Skandalis, S.S.; Neill, T.; Multhaupt, H.A.; Hubo, M.; Frey, H.; Gopal, S.; Gomes, A.; Afratis, N.; Lim, H.C.; et al. Insights into the key roles of proteoglycans in breast cancer biology and translational medicine. Biochim. Biophys. Acta 2015, 1855, 276–300. [Google Scholar] [CrossRef] [PubMed]
  116. Sanderson, R.D.; Yang, Y.; Suva, L.J.; Kelly, T. Heparan sulfate proteoglycans and heparanase—Partners in osteolytic tumor growth and metastasis. Matrix Biol. 2004, 23, 341–352. [Google Scholar] [CrossRef] [PubMed]
  117. Ritchie, J.P.; Ramani, V.C.; Ren, Y.; Naggi, A.; Torri, G.; Casu, B.; Penco, S.; Pisano, C.; Carminati, P.; Tortoreto, M.; et al. SST0001, a chemically modified heparin, inhibits myeloma growth and angiogenesis via disruption of the heparanase/syndecan-1 axis. Clin. Cancer Res. 2011, 17, 1382–1393. [Google Scholar] [CrossRef] [PubMed]
  118. Filmus, J.; Selleck, S.B. Glypicans: Proteoglycans with a surprise. J. Clin. Investig. 2001, 108, 497–501. [Google Scholar] [CrossRef] [PubMed]
  119. Suhovskih, A.V.; Aidagulova, S.V.; Kashuba, V.I.; Grigorieva, E.V. Proteoglycans as potential microenvironmental biomarkers for colon cancer. Cell Tissue Res. 2015, 361, 833–844. [Google Scholar] [CrossRef] [PubMed]
  120. Baghy, K.; Tatrai, P.; Regos, E.; Kovalszky, I. Proteoglycans in liver cancer. World J. Gastroenterol. 2016, 22, 379–393. [Google Scholar] [CrossRef] [PubMed]
  121. Murdoch, A.D.; Dodge, G.R.; Cohen, I.; Tuan, R.S.; Iozzo, R.V. Primary structure of the human heparan sulfate proteoglycan from basement membrane (HSPG2/perlecan). A chimeric molecule with multiple domains homologous to the low density lipoprotein receptor, laminin, neural cell adhesion molecules, and epidermal growth factor. J. Biol. Chem. 1992, 267, 8544–8557. [Google Scholar] [PubMed]
  122. Iozzo, R.V.; Cohen, I.R.; Grassel, S.; Murdoch, A.D. The biology of perlecan: The multifaceted heparan sulphate proteoglycan of basement membranes and pericellular matrices. Biochem. J. 1994, 302, 625–639. [Google Scholar] [CrossRef] [PubMed]
  123. Farach-Carson, M.C.; Carson, D.D. Perlecan-a multifunctional extracellular proteoglycan scaffold. Glycobiology 2007, 17, 897–905. [Google Scholar] [CrossRef] [PubMed]
  124. Murdoch, A.D.; Liu, B.; Schwarting, R.; Tuan, R.S.; Iozzo, R.V. Widespread expression of perlecan proteoglycan in basement membranes and extracellular matrices of human tissues as detected by a novel monoclonal antibody against domain III and by in situ hybridization. J. Histochem. Cytochem. 1994, 42, 239–249. [Google Scholar] [CrossRef] [PubMed]
  125. Handler, M.; Yurchenco, P.D.; Iozzo, R.V. Developmental expression of perlecan during murine embryogenesis. Dev. Dyn. 1997, 210, 130–145. [Google Scholar] [CrossRef]
  126. Whitelock, J.M.; Melrose, J.; Iozzo, R.V. Diverse cell signaling events modulated by perlecan. Biochemistry 2008, 47, 11174–11183. [Google Scholar] [CrossRef] [PubMed]
  127. Mongiat, M.; Taylor, K.; Otto, J.; Aho, S.; Uitto, J.; Whitelock, J.M.; Iozzo, R.V. The protein core of the proteoglycan perlecan binds specifically to fibroblast growth factor-7. J. Biol. Chem. 2000, 275, 7095–7100. [Google Scholar] [CrossRef] [PubMed]
  128. Mongiat, M.; Otto, J.; Oldershaw, R.; Ferrer, F.; Sato, J.D.; Iozzo, R.V. Fibroblast growth factor-binding protein is a novel partner for perlecan protein core. J. Biol. Chem. 2001, 276, 10263–10271. [Google Scholar] [CrossRef] [PubMed]
  129. Mongiat, M.; Fu, J.; Oldershaw, R.; Greenhalgh, R.; Gown, A.M.; Iozzo, R.V. Perlecan protein core interacts with extracellular matrix protein 1 (ECM1), a glycoprotein involved in bone formation and angiogenesis. J. Biol. Chem. 2003, 278, 17491–17499. [Google Scholar] [CrossRef] [PubMed]
  130. Gonzalez, E.M.; Mongiat, M.; Slater, S.J.; Baffa, R.; Iozzo, R.V. A novel interaction between perlecan protein core and progranulin: Potential effects on tumor growth. J. Biol. Chem. 2003, 278, 38113–38116. [Google Scholar] [CrossRef] [PubMed]
  131. Li, S.; Shimono, C.; Norioka, N.; Nakano, I.; Okubo, T.; Yagi, Y.; Hayashi, M.; Sato, Y.; Fujisaki, H.; Hattori, S.; et al. Activin A binds to perlecan through its pro-region that has heparin/heparan sulfate binding activity. J. Biol. Chem. 2010, 285, 36645–36655. [Google Scholar] [CrossRef] [PubMed]
  132. Poluzzi, C.; Iozzo, R.V.; Schaefer, L. Endostatin and endorepellin: A common route of action for similar angiostatic cancer avengers. Adv. Drug Deliv. Rev. 2016, 97, 156–173. [Google Scholar] [CrossRef] [PubMed]
  133. Sharma, B.; Handler, M.; Eichstetter, I.; Whitelock, J.M.; Nugent, M.A.; Iozzo, R.V. Antisense targeting of perlecan blocks tumor growth and angiogenesis in vivo. J. Clin. Investig. 1998, 102, 1599–1608. [Google Scholar] [CrossRef] [PubMed]
  134. Iozzo, R.V.; San Antonio, J.D. Heparan sulfate proteoglycans: Heavy hitters in the angiogenesis arena. J. Clin. Investig. 2001, 108, 349–355. [Google Scholar] [CrossRef] [PubMed]
  135. Zhou, Z.; Wang, J.; Cao, R.; Morita, H.; Soininen, R.; Chan, K.M.; Liu, B.; Cao, Y.; Tryggvason, K. Impaired angiogenesis, delayed wound healing and retarded tumor growth in perlecan heparan sulfate-deficient mice. Cancer Res. 2004, 64, 4699–4702. [Google Scholar] [CrossRef] [PubMed]
  136. Aviezer, D.; Hecht, D.; Safran, M.; Eisinger, M.; David, G.; Yayon, A. Perlecan, basal lamina proteoglycan, promotes basic fibroblast growth factor-receptor binding, mitogenesis, and angiogenesis. Cell 1994, 79, 1005–1013. [Google Scholar] [CrossRef]
  137. Ghiselli, G.; Eichstetter, I.; Iozzo, R.V. A role for the perlecan protein core in the activation of the keratinocyte growth factor receptor. Biochem. J. 2001, 359, 153–163. [Google Scholar] [CrossRef] [PubMed]
  138. Murakami, M.; Simons, M. Fibroblast growth factor regulation of neovascularization. Curr. Opin. Hematol. 2008, 15, 215–220. [Google Scholar] [CrossRef] [PubMed]
  139. Zoeller, J.J.; Whitelock, J.M.; Iozzo, R.V. Perlecan regulates developmental angiogenesis by modulating the VEGF-VEGFR2 axis. Matrix Biol. 2009, 28, 284–291. [Google Scholar] [CrossRef] [PubMed]
  140. Chuang, C.Y.; Lord, M.S.; Melrose, J.; Rees, M.D.; Knox, S.M.; Freeman, C.; Iozzo, R.V.; Whitelock, J.M. Heparan sulfate-dependent signaling of fibroblast growth factor 18 by chondrocyte-derived perlecan. Biochemistry 2010, 49, 5524–5532. [Google Scholar] [CrossRef] [PubMed]
  141. Muthusamy, A.; Cooper, C.R.; Gomes, R.R., Jr. Soluble perlecan domain I enhances vascular endothelial growth factor-165 activity and receptor phosphorylation in human bone marrow endothelial cells. BMC Biochem. 2010, 11, 43. [Google Scholar] [CrossRef] [PubMed]
  142. Lord, M.S.; Chuang, C.Y.; Melrose, J.; Davies, M.J.; Iozzo, R.V.; Whitelock, J.M. The role of vascular-derived perlecan in modulating cell adhesion, proliferation and growth factor signaling. Matrix Biol. 2014, 35, 112–122. [Google Scholar] [CrossRef] [PubMed]
  143. Whitelock, J.M.; Murdoch, A.D.; Iozzo, R.V.; Underwood, P.A. The degradation of human endothelial cell-derived perlecan and release of bound basic fibroblast growth factor by stromelysin, collagenase, plasmin, and heparanases. J. Biol. Chem. 1996, 271, 10079–10086. [Google Scholar] [PubMed]
  144. Marchisone, C.; Del Grosso, F.; Masiello, L.; Prat, M.; Santi, L.; Noonan, D.M. Phenotypic alterations in Kaposi's sarcoma cells by antisense reduction of perlecan. Pathol. Oncol. Res. 2000, 6, 10–17. [Google Scholar] [CrossRef] [PubMed]
  145. Chang, J.W.; Kang, U.B.; Kim, D.H.; Yi, J.K.; Lee, J.W.; Noh, D.Y.; Lee, C.; Yu, M.H. Identification of circulating endorepellin LG3 fragment: Potential use as a serological biomarker for breast cancer. Proteom. Clin. Appl. 2008, 2, 23–32. [Google Scholar] [CrossRef] [PubMed]
  146. Kawahara, R.; Granato, D.C.; Carnielli, C.M.; Cervigne, N.K.; Oliveria, C.E.; Rivera, C.; Yokoo, S.; Fonseca, F.P.; Lopes, M.; Santos-Silva, A.R.; et al. Agrin and perlecan mediate tumorigenic processes in oral squamous cell carcinoma. PLoS ONE 2014, 9, e115004. [Google Scholar] [CrossRef] [PubMed]
  147. Grindel, B.; Li, Q.; Arnold, R.; Petros, J.; Zayzafoon, M.; Muldoon, M.; Stave, J.; Chung, L.W.; Farach-Carson, M.C. Perlecan/HSPG2 and matrilysin/MMP-7 as indices of tissue invasion: Tissue localization and circulating perlecan fragments in a cohort of 288 radical prostatectomy patients. Oncotarget 2016, 7, 10433–10447. [Google Scholar] [CrossRef] [PubMed]
  148. Mongiat, M.; Sweeney, S.M.; San Antonio, J.D.; Fu, J.; Iozzo, R.V. Endorepellin, a novel inhibitor of angiogenesis derived from the C terminus of perlecan. J. Biol. Chem. 2003, 278, 4238–4249. [Google Scholar] [CrossRef] [PubMed]
  149. Bix, G.; Fu, J.; Gonzalez, E.M.; Macro, L.; Barker, A.; Campbell, S.; Zutter, M.M.; Santoro, S.A.; Kim, J.K.; Hook, M.; et al. Endorepellin causes endothelial cell disassembly of actin cytoskeleton and focal adhesions through α2β1 integrin. J. Cell Biol. 2004, 166, 97–109. [Google Scholar] [CrossRef] [PubMed]
  150. Bix, G.; Castello, R.; Burrows, M.; Zoeller, J.J.; Weech, M.; Iozzo, R.A.; Cardi, C.; Thakur, M.L.; Barker, C.A.; Camphausen, K.; et al. Endorepellin in vivo: Targeting the tumor vasculature and retarding cancer growth and metabolism. J. Natl. Cancer Inst. 2006, 98, 1634–1646. [Google Scholar] [CrossRef] [PubMed]
  151. Woodall, B.P.; Nystrom, A.; Iozzo, R.A.; Eble, J.A.; Niland, S.; Krieg, T.; Eckes, B.; Pozzi, A.; Iozzo, R.V. Integrin α2β1 is the required receptor for endorepellin angiostatic activity. J. Biol. Chem. 2008, 283, 2335–2343. [Google Scholar] [CrossRef] [PubMed]
  152. Nystrom, A.; Shaik, Z.P.; Gullberg, D.; Krieg, T.; Eckes, B.; Zent, R.; Pozzi, A.; Iozzo, R.V. Role of tyrosine phosphatase SHP-1 in the mechanism of endorepellin angiostatic activity. Blood 2009, 114, 4897–4906. [Google Scholar] [CrossRef] [PubMed]
  153. Goyal, A.; Poluzzi, C.; Willis, C.D.; Smythies, J.; Shellard, A.; Neill, T.; Iozzo, R.V. Endorepellin affects angiogenesis by antagonizing diverse vascular endothelial growth factor receptor 2 (VEGFR2)-evoked signaling pathways: Transcriptional repression of hypoxia-inducible factor 1α and VEGFA and concurrent inhibition of nuclear factor of activated T cell 1 (NFAT1) activation. J. Biol. Chem. 2012, 287, 43543–43556. [Google Scholar] [PubMed]
  154. Goyal, A.; Pal, N.; Concannon, M.; Paul, M.; Doran, M.; Poluzzi, C.; Sekiguchi, K.; Whitelock, J.M.; Neill, T.; Iozzo, R.V. Endorepellin, the angiostatic module of perlecan, interacts with both the α2β1 integrin and vascular endothelial growth factor receptor 2 (VEGFR2): A dual receptor antagonism. J. Biol. Chem. 2011, 286, 25947–25962. [Google Scholar] [CrossRef] [PubMed]
  155. Douglass, S.; Goyal, A.; Iozzo, R.V. The role of perlecan and endorepellin in the control of tumor angiogenesis and endothelial cell autophagy. Connect. Tissue Res. 2015, 56, 381–391. [Google Scholar] [CrossRef] [PubMed]
  156. Poluzzi, C.; Casulli, J.; Goyal, A.; Mercer, T.J.; Neill, T.; Iozzo, R.V. Endorepellin evokes autophagy in endothelial cells. J. Biol. Chem. 2014, 289, 16114–16128. [Google Scholar] [CrossRef] [PubMed]
  157. Goyal, A.; Gubbiotti, M.A.; Chery, D.R.; Han, L.; Iozzo, R.V. Endorepellin-evoked autophagy contributes to angiostasis. J. Biol. Chem. 2016. [Google Scholar] [CrossRef] [PubMed]
  158. Sofeu Feugaing, D.D.; Gotte, M.; Viola, M. More than matrix: The multifaceted role of decorin in cancer. Eur. J. Cell Biol. 2013, 92, 1–11. [Google Scholar] [CrossRef] [PubMed]
  159. Jarvelainen, H.; Vernon, R.B.; Gooden, M.D.; Francki, A.; Lara, S.; Johnson, P.Y.; Kinsella, M.G.; Sage, E.H.; Wight, T.N. Overexpression of decorin by rat arterial smooth muscle cells enhances contraction of type I collagen in vitro. Arterioscler. Thromb. Vasc. Biol. 2004, 24, 67–72. [Google Scholar] [CrossRef] [PubMed]
  160. Kalamajski, S.; Oldberg, A. The role of small leucine-rich proteoglycans in collagen fibrillogenesis. Matrix Biol. 2010, 29, 248–253. [Google Scholar] [CrossRef] [PubMed]
  161. Reese, S.P.; Underwood, C.J.; Weiss, J.A. Effects of decorin proteoglycan on fibrillogenesis, ultrastructure, and mechanics of type I collagen gels. Matrix Biol. 2013, 32, 414–423. [Google Scholar] [CrossRef] [PubMed]
  162. Morcavallo, A.; Buraschi, S.; Xu, S.Q.; Belfiore, A.; Schaefer, L.; Iozzo, R.V.; Morrione, A. Decorin differentially modulates the activity of insulin receptor isoform A ligands. Matrix Biol. 2014, 35, 82–90. [Google Scholar] [CrossRef] [PubMed]
  163. Mohan, R.R.; Tovey, J.C.; Sharma, A.; Schultz, G.S.; Cowden, J.W.; Tandon, A. Targeted decorin gene therapy delivered with adeno-associated virus effectively retards corneal neovascularization in vivo. PLoS ONE 2011, 6, e26432. [Google Scholar] [CrossRef] [PubMed]
  164. Scott, R.A.; Panitch, A. Decorin mimic regulates platelet-derived growth factor and interferon-gamma stimulation of vascular smooth muscle cells. Biomacromolecules 2014, 15, 2090–2103. [Google Scholar] [CrossRef] [PubMed]
  165. Jarvelainen, H.; Sainio, A.; Wight, T.N. Pivotal role for decorin in angiogenesis. Matrix Biol. 2015, 43, 15–26. [Google Scholar] [CrossRef] [PubMed]
  166. Imai, K.; Hiramatsu, A.; Fukushima, D.; Pierschbacher, M.D.; Okada, Y. Degradation of decorin by matrix metalloproteinases: Identification of the cleavage sites, kinetic analyses and transforming growth factor-β1 release. Biochem. J. 1997, 322, 809–814. [Google Scholar] [CrossRef] [PubMed]
  167. Boivin, W.A.; Shackleford, M.; Vanden Hoek, A.; Zhao, H.; Hackett, T.L.; Knight, D.A.; Granville, D.J. Granzyme B cleaves decorin, biglycan and soluble βglycan, releasing active transforming growth factor-β1. PLoS ONE 2012, 7, e33163. [Google Scholar] [CrossRef]
  168. Neill, T.; Schaefer, L.; Iozzo, R.V. Decorin as a multivalent therapeutic agent against cancer. Adv. Drug Deliv. Rev. 2016, 97, 174–185. [Google Scholar] [CrossRef] [PubMed]
  169. Neill, T.; Painter, H.; Buraschi, S.; Owens, R.T.; Lisanti, M.P.; Schaefer, L.; Iozzo, R.V. Decorin antagonizes the angiogenic network: Concurrent inhibition of Met, hypoxia inducible factor 1α, vascular endothelial growth factor A, and induction of thrombospondin-1 and TIMP3. J. Biol. Chem. 2012, 287, 5492–5506. [Google Scholar] [CrossRef] [PubMed]
  170. Goyal, A.; Neill, T.; Owens, R.T.; Schaefer, L.; Iozzo, R.V. Decorin activates AMPK, an energy sensor kinase, to induce autophagy in endothelial cells. Matrix Biol. 2014, 35, 42–50. [Google Scholar] [CrossRef] [PubMed]
  171. Buraschi, S.; Neill, T.; Goyal, A.; Poluzzi, C.; Smythies, J.; Owens, R.T.; Schaefer, L.; Torres, A.; Iozzo, R.V. Decorin causes autophagy in endothelial cells via Peg3. Proc. Natl. Acad. Sci. USA 2013, 110, E2582–E2591. [Google Scholar] [CrossRef] [PubMed]
  172. Gubbiotti, M.A.; Iozzo, R.V. Proteoglycans regulate autophagy via outside-in signaling: An emerging new concept. Matrix Biol. 2015, 48, 6–13. [Google Scholar] [CrossRef] [PubMed]
  173. Schonherr, E.; Sunderkotter, C.; Schaefer, L.; Thanos, S.; Grassel, S.; Oldberg, A.; Iozzo, R.V.; Young, M.F.; Kresse, H. Decorin deficiency leads to impaired angiogenesis in injured mouse cornea. J. Vasc. Res. 2004, 41, 499–508. [Google Scholar] [CrossRef] [PubMed]
  174. Jarvelainen, H.; Puolakkainen, P.; Pakkanen, S.; Brown, E.L.; Hook, M.; Iozzo, R.V.; Sage, E.H.; Wight, T.N. A role for decorin in cutaneous wound healing and angiogenesis. Wound Repair Regen. 2006, 14, 443–452. [Google Scholar] [CrossRef] [PubMed]
  175. Grant, D.S.; Yenisey, C.; Rose, R.W.; Tootell, M.; Santra, M.; Iozzo, R.V. Decorin suppresses tumor cell-mediated angiogenesis. Oncogene 2002, 21, 4765–4777. [Google Scholar] [CrossRef] [PubMed]
  176. Berendsen, A.D.; Pinnow, E.L.; Maeda, A.; Brown, A.C.; McCartney-Francis, N.; Kram, V.; Owens, R.T.; Robey, P.G.; Holmbeck, K.; de Castro, L.F.; et al. Biglycan modulates angiogenesis and bone formation during fracture healing. Matrix Biol. 2014, 35, 223–231. [Google Scholar] [CrossRef] [PubMed]
  177. Myren, M.; Kirby, D.J.; Noonan, M.L.; Maeda, A.; Owens, R.T.; Ricard-Blum, S.; Kram, V.; Kilts, T.M.; Young, M.F. Biglycan potentially regulates angiogenesis during fracture repair by altering expression and function of endostatin. Matrix Biol. 2016, 52–54, 141–150. [Google Scholar] [CrossRef] [PubMed]
  178. Xing, X.; Gu, X.; Ma, T.; Ye, H. Biglycan up-regulated vascular endothelial growth factor (VEGF) expression and promoted angiogenesis in colon cancer. Tumour Biol. 2015, 36, 1773–1780. [Google Scholar] [CrossRef] [PubMed]
  179. Barbouri, D.; Afratis, N.; Gialeli, C.; Vynios, D.H.; Theocharis, A.D.; Karamanos, N.K. Syndecans as modulators and potential pharmacological targets in cancer progression. Front. Oncol. 2014, 4, 4. [Google Scholar] [CrossRef] [PubMed]
  180. Rapraeger, A.C.; Ell, B.J.; Roy, M.; Li, X.; Morrison, O.R.; Thomas, G.M.; Beauvais, D.M. Vascular endothelial-cadherin stimulates syndecan-1-coupled insulin-like growth factor-1 receptor and cross-talk between αvβ3 integrin and vascular endothelial growth factor receptor 2 at the onset of endothelial cell dissemination during angiogenesis. FEBS J. 2013, 280, 2194–2206. [Google Scholar] [CrossRef] [PubMed]
  181. Lamorte, S.; Ferrero, S.; Aschero, S.; Monitillo, L.; Bussolati, B.; Omede, P.; Ladetto, M.; Camussi, G. Syndecan-1 promotes the angiogenic phenotype of multiple myeloma endothelial cells. Leukemia 2012, 26, 1081–1090. [Google Scholar] [CrossRef] [PubMed]
  182. Jing, Z.; Wei-Jie, Y.; Yi-Feng, Z.G.; Jing, H. Downregulation of Syndecan-1 induce glomerular endothelial cell dysfunction through modulating internalization of VEGFR-2. Cell Signal. 2016, 28, 826–837. [Google Scholar] [CrossRef] [PubMed]
  183. Maeda, T.; Desouky, J.; Friedl, A. Syndecan-1 expression by stromal fibroblasts promotes breast carcinoma growth in vivo and stimulates tumor angiogenesis. Oncogene 2006, 25, 1408–1412. [Google Scholar] [CrossRef] [PubMed]
  184. Noguer, O.; Villena, J.; Lorita, J.; Vilaro, S.; Reina, M. Syndecan-2 downregulation impairs angiogenesis in human microvascular endothelial cells. Exp. Cell Res. 2009, 315, 795–808. [Google Scholar] [CrossRef] [PubMed]
  185. Purushothaman, A.; Uyama, T.; Kobayashi, F.; Yamada, S.; Sugahara, K.; Rapraeger, A.C.; Sanderson, R.D. Heparanase-enhanced shedding of syndecan-1 by myeloma cells promotes endothelial invasion and angiogenesis. Blood 2010, 115, 2449–2457. [Google Scholar] [CrossRef] [PubMed]
  186. De Rossi, G.; Evans, A.R.; Kay, E.; Woodfin, A.; McKay, T.R.; Nourshargh, S.; Whiteford, J.R. Shed syndecan-2 inhibits angiogenesis. J. Cell Sci. 2014, 127, 4788–4799. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Rodriguez-Manzaneque, J.C.; Carpizo, D.; Plaza-Calonge, M.C.; Torres-Collado, A.X.; Thai, S.N.; Simons, M.; Horowitz, A.; Iruela-Arispe, M.L. Cleavage of syndecan-4 by ADAMTS1 provokes defects in adhesion. Int. J. Biochem. Cell Biol. 2009, 41, 800–810. [Google Scholar] [CrossRef] [PubMed]
  188. Li, R.; Xie, J.; Wu, H.; Li, G.; Chen, J.; Chen, Q.; Wang, L.; Xu, B. Syndecan-4 shedding impairs macrovascular angiogenesis in diabetes mellitus. Biochem. Biophys. Res. Commun. 2016, 474, 15–21. [Google Scholar] [CrossRef] [PubMed]
  189. Aikawa, T.; Whipple, C.A.; Lopez, M.E.; Gunn, J.; Young, A.; Lander, A.D.; Korc, M. Glypican-1 modulates the angiogenic and metastatic potential of human and mouse cancer cells. J. Clin. Investig. 2008, 118, 89–99. [Google Scholar] [CrossRef] [PubMed]
  190. Qiao, D.; Meyer, K.; Mundhenke, C.; Drew, S.A.; Friedl, A. Heparan sulfate proteoglycans as regulators of fibroblast growth factor-2 signaling in brain endothelial cells. Specific role for glypican-1 in glioma angiogenesis. J. Biol. Chem. 2003, 278, 16045–16053. [Google Scholar] [CrossRef] [PubMed]
  191. Fico, A.; Maina, F.; Dono, R. Fine-tuning of cell signaling by glypicans. Cell. Mol. Life Sci. 2011, 68, 923–929. [Google Scholar] [CrossRef] [PubMed]
  192. Monteforte, A.J.; Lam, B.; Das, S.; Mukhopadhyay, S.; Wright, C.S.; Martin, P.E.; Dunn, A.K.; Baker, A.B. Glypican-1 nanoliposomes for potentiating growth factor activity in therapeutic angiogenesis. Biomaterials 2016, 94, 45–56. [Google Scholar] [CrossRef] [PubMed]
  193. Naito, Z. Role of the small leucine-rich proteoglycan (SLRP) family in pathological lesions and cancer cell growth. J. Nippon Med. Sch. 2005, 72, 137–145. [Google Scholar] [CrossRef] [PubMed]
  194. D'Onofrio, M.F.; Brezillon, S.; Baranek, T.; Perreau, C.; Roughley, P.J.; Maquart, F.X.; Wegrowski, Y. Identification of β1 integrin as mediator of melanoma cell adhesion to lumican. Biochem. Biophys. Res. Commun. 2008, 365, 266–272. [Google Scholar] [CrossRef] [PubMed]
  195. Albig, A.R.; Roy, T.G.; Becenti, D.J.; Schiemann, W.P. Transcriptome analysis of endothelial cell gene expression induced by growth on matrigel matrices: Identification and characterization of MAGP-2 and lumican as novel regulators of angiogenesis. Angiogenesis 2007, 10, 197–216. [Google Scholar] [CrossRef] [PubMed]
  196. Niewiarowska, J.; Brezillon, S.; Sacewicz-Hofman, I.; Bednarek, R.; Maquart, F.X.; Malinowski, M.; Wiktorska, M.; Wegrowski, Y.; Cierniewski, C.S. Lumican inhibits angiogenesis by interfering with α2β1 receptor activity and downregulating MMP-14 expression. Thromb. Res. 2011, 128, 452–457. [Google Scholar] [CrossRef] [PubMed]
  197. Nikitovic, D.; Papoutsidakis, A.; Karamanos, N.K.; Tzanakakis, G.N. Lumican affects tumor cell functions, tumor-ECM interactions, angiogenesis and inflammatory response. Matrix Biol. 2014, 35, 206–214. [Google Scholar] [CrossRef] [PubMed]
  198. Williams, K.E.; Fulford, L.A.; Albig, A.R. Lumican reduces tumor growth via induction of fas-mediated endothelial cell apoptosis. Cancer Microenviron. 2010, 4, 115–126. [Google Scholar] [CrossRef] [PubMed]
  199. Chanmee, T.; Ontong, P.; Itano, N. Hyaluronan: A modulator of the tumor microenvironment. Cancer Lett. 2016, 375, 20–30. [Google Scholar] [CrossRef] [PubMed]
  200. Slevin, M.; Kumar, S.; Gaffney, J. Angiogenic oligosaccharides of hyaluronan induce multiple signaling pathways affecting vascular endothelial cell mitogenic and wound healing responses. J. Biol. Chem. 2002, 277, 41046–41059. [Google Scholar] [CrossRef] [PubMed]
  201. Toole, B.P. Hyaluronan: From extracellular glue to pericellular cue. Nat. Rev. Cancer 2004, 4, 528–539. [Google Scholar] [CrossRef] [PubMed]
  202. Savani, R.C.; Cao, G.; Pooler, P.M.; Zaman, A.; Zhou, Z.; DeLisser, H.M. Differential involvement of the hyaluronan (HA) receptors CD44 and receptor for HA-mediated motility in endothelial cell function and angiogenesis. J. Biol. Chem. 2001, 276, 36770–36778. [Google Scholar] [CrossRef] [PubMed]
  203. Slevin, M.; Krupinski, J.; Gaffney, J.; Matou, S.; West, D.; Delisser, H.; Savani, R.C.; Kumar, S. Hyaluronan-mediated angiogenesis in vascular disease: Uncovering RHAMM and CD44 receptor signaling pathways. Matrix Biol. 2007, 26, 58–68. [Google Scholar] [CrossRef] [PubMed]
  204. Gao, F.; Yang, C.X.; Mo, W.; Liu, Y.W.; He, Y.Q. Hyaluronan oligosaccharides are potential stimulators to angiogenesis via RHAMM mediated signal pathway in wound healing. Clin. Investig. Med. 2008, 31, E106–E116. [Google Scholar]
  205. Wang, Y.; Han, G.; Guo, B.; Huang, J. Hyaluronan oligosaccharides promote diabetic wound healing by increasing angiogenesis. Pharmacol. Rep. 2016, 68, 1126–1132. [Google Scholar] [CrossRef] [PubMed]
  206. Olivares, C.N.; Alaniz, L.D.; Menger, M.D.; Baranao, R.I.; Laschke, M.W.; Meresman, G.F. Inhibition of Hyaluronic Acid Synthesis Suppresses Angiogenesis in Developing Endometriotic Lesions. PLoS ONE 2016, 11, e0152302. [Google Scholar] [CrossRef] [PubMed]
  207. Litwiniuk, M.; Krejner, A.; Speyrer, M.S.; Gauto, A.R.; Grzela, T. Hyaluronic Acid in Inflammation and Tissue Regeneration. Wounds 2016, 28, 78–88. [Google Scholar] [PubMed]
  208. West, D.C.; Kumar, S. The effect of hyaluronate and its oligosaccharides on endothelial cell proliferation and monolayer integrity. Exp. Cell Res. 1989, 183, 179–196. [Google Scholar] [CrossRef]
  209. Bollyky, P.L.; Lord, J.D.; Masewicz, S.A.; Evanko, S.P.; Buckner, J.H.; Wight, T.N.; Nepom, G.T. Cutting edge: High molecular weight hyaluronan promotes the suppressive effects of CD4+CD25+ regulatory T cells. J. Immunol. 2007, 179, 744–747. [Google Scholar] [CrossRef] [PubMed]
  210. Fuchs, K.; Hippe, A.; Schmaus, A.; Homey, B.; Sleeman, J.P.; Orian-Rousseau, V. Opposing effects of high- and low-molecular weight hyaluronan on CXCL12-induced CXCR4 signaling depend on CD44. Cell Death Dis. 2013, 4, e819. [Google Scholar] [CrossRef] [PubMed]
  211. Koyama, H.; Hibi, T.; Isogai, Z.; Yoneda, M.; Fujimori, M.; Amano, J.; Kawakubo, M.; Kannagi, R.; Kimata, K.; Taniguchi, S.; et al. Hyperproduction of hyaluronan in neu-induced mammary tumor accelerates angiogenesis through stromal cell recruitment: Possible involvement of versican/PG-M. Am. J. Pathol. 2007, 170, 1086–1099. [Google Scholar] [CrossRef] [PubMed]
  212. Singleton, P.A. Hyaluronan regulation of endothelial barrier function in cancer. Adv. Cancer Res. 2014, 123, 191–209. [Google Scholar] [PubMed]
  213. Doliana, R.; Canton, A.; Bucciotti, F.; Mongiat, M.; Bonaldo, P.; Colombatti, A. Structure, chromosomal localization, and promoter analysis of the human elastin microfibril interfase located proteIN (EMILIN) gene. J. Biol. Chem. 2000, 275, 785–792. [Google Scholar] [CrossRef] [PubMed]
  214. Mongiat, M.; Mungiguerra, G.; Bot, S.; Mucignat, M.T.; Giacomello, E.; Doliana, R.; Colombatti, A. Self-assembly and supramolecular organization of EMILIN. J. Biol. Chem. 2000, 275, 25471–25480. [Google Scholar] [CrossRef] [PubMed]
  215. Colombatti, A.; Spessotto, P.; Doliana, R.; Mongiat, M.; Bressan, G.M.; Esposito, G. The EMILIN/Multimerin family. Front. Immunol. 2011, 2, 93. [Google Scholar] [CrossRef] [PubMed]
  216. Bot, S.; Andreuzzi, E.; Capuano, A.; Schiavinato, A.; Colombatti, A.; Doliana, R. Multiple-interactions among EMILIN1 and EMILI. Matrix Biol. 2015, 41, 44–55. [Google Scholar] [CrossRef] [PubMed]
  217. Jeimy, S.B.; Tasneem, S.; Cramer, E.M.; Hayward, C.P. Multimerin 1. Platelets 2008, 19, 83–95. [Google Scholar] [CrossRef] [PubMed]
  218. Christian, S.; Ahorn, H.; Novatchkova, M.; Garin-Chesa, P.; Park, J.E.; Weber, G.; Eisenhaber, F.; Rettig, W.J.; Lenter, M.C. Molecular cloning and characterization of EndoGlyx-1, an EMILIN-like multisubunit glycoprotein of vascular endothelium. J. Biol. Chem. 2001, 276, 48588–48595. [Google Scholar] [PubMed]
  219. Doliana, R.; Mongiat, M.; Bucciotti, F.; Giacomello, E.; Deutzmann, R.; Volpin, D.; Bressan, G.M.; Colombatti, A. EMILIN, a component of the elastic fiber and a new member of the C1q/tumor necrosis factor superfamily of proteins. J. Biol. Chem. 1999, 274, 16773–16781. [Google Scholar] [CrossRef] [PubMed]
  220. Doliana, R.; Bot, S.; Mungiguerra, G.; Canton, A.; Cilli, S.P.; Colombatti, A. Isolation and characterization of EMILIN-2, a new component of the growing EMILINs family and a member of the EMI domain-containing superfamily. J. Biol. Chem. 2001, 276, 12003–12011. [Google Scholar] [CrossRef] [PubMed]
  221. Corallo, D.; Schiavinato, A.; Trapani, V.; Moro, E.; Argenton, F.; Bonaldo, P. Emilin3 is required for notochord sheath integrity and interacts with Scube2 to regulate notochord-derived Hedgehog signals. Development 2013, 140, 4594–4601. [Google Scholar] [CrossRef] [PubMed]
  222. Leimeister, C.; Steidl, C.; Schumacher, N.; Erhard, S.; Gessler, M. Developmental Expression and Biochemical Characterization of Emu Family Members. Dev. Biol. 2002, 249, 204–218. [Google Scholar] [CrossRef] [PubMed]
  223. Sanz-Moncasi, M.P.; Garin-Chesa, P.; Stockert, E.; Jaffe, E.A.; Old, L.J.; Rettig, W.J. Identification of a high molecular weight endothelial cell surface glycoprotein, endoGlyx-1, in normal and tumor blood vessels. Lab. Investig. 1994, 71, 366–373. [Google Scholar] [PubMed]
  224. Lorenzon, E.; Colladel, R.; Andreuzzi, E.; Marastoni, S.; Todaro, F.; Schiappacassi, M.; Ligresti, G.; Colombatti, A.; Mongiat, M. MULTIMERIN2 impairs tumor angiogenesis and growth by interfering with VEGF-A/VEGFR2 pathway. Oncogene 2012, 31, 3136–3147. [Google Scholar] [CrossRef] [PubMed]
  225. Colladel, R.; Pellicani, R.; Andreuzzi, E.; Paulitti, A.; Tarticchio, G.; Todaro, F.; Colombatti, A.; Mongiat, M. MULTIMERIN2 binds VEGF-A primarily via the carbohydrate chains exerting an angiostatic function and impairing tumor growth. Oncotarget 2016, 7, 2022–2037. [Google Scholar] [PubMed]
  226. Noy, P.J.; Swain, R.K.; Khan, K.; Lodhia, P.; Bicknell, R. Sprouting angiogenesis is regulated by shedding of the C-type lectin family 14, member A (CLEC14A) ectodomain, catalyzed by rhomboid-like 2 protein (RHBDL2). FASEB J. 2016, 30, 2311–2323. [Google Scholar] [CrossRef] [PubMed]
  227. Noy, P.J.; Lodhia, P.; Khan, K.; Zhuang, X.; Ward, D.G.; Verissimo, A.R.; Bacon, A.; Bicknell, R. Blocking CLEC14A-MMRN2 binding inhibits sprouting angiogenesis and tumour growth. Oncogene 2015, 34, 5821–5831. [Google Scholar] [CrossRef] [PubMed]
  228. Braghetta, P.; Ferrari, A.; de Gemmis, P.; Zanetti, M.; Volpin, D.; Bonaldo, P.; Bressan, G.M. Overlapping, complementary and site-specific expression pattern of genes of the EMILIN/Multimerin family. Matrix Biol. 2004, 22, 549–556. [Google Scholar] [CrossRef] [PubMed]
  229. Hill, V.K.; Hesson, L.B.; Dansranjavin, T.; Dallol, A.; Bieche, I.; Vacher, S.; Tommasi, S.; Dobbins, T.; Gentle, D.; Euhus, D.; et al. Identification of 5 novel genes methylated in breast and other epithelial cancers. Mol. Cancer 2010, 9, 51. [Google Scholar] [CrossRef] [PubMed]
  230. Mongiat, M.; Marastoni, S.; Ligresti, G.; Lorenzon, E.; Schiappacassi, M.; Perris, R.; Frustaci, S.; Colombatti, A. The extracellular matrix glycoprotein elastin microfibril interface located protein 2: A dual role in the tumor microenvironment. Neoplasia 2010, 12, 294–304. [Google Scholar] [CrossRef] [PubMed]
  231. Bronisz, A.; Godlewski, J.; Wallace, J.A.; Merchant, A.S.; Nowicki, M.O.; Mathsyaraja, H.; Srinivasan, R.; Trimboli, A.J.; Martin, C.K.; Li, F.; et al. Reprogramming of the tumour microenvironment by stromal PTEN-regulated miR-320. Nat. Cell Biol. 2012, 14, 159–167. [Google Scholar] [CrossRef] [PubMed]
  232. Klenotic, P.A.; Zhang, C.; Lin, Z. Emerging roles of CCN proteins in vascular development and pathology. J. Cell Commun. Signal. 2016, 10, 251–257. [Google Scholar] [CrossRef] [PubMed]
  233. Jun, J.I.; Lau, L.F. Taking aim at the extracellular matrix: CCN proteins as emerging therapeutic targets. Nat. Rev. Drug Discov. 2011, 10, 945–963. [Google Scholar] [CrossRef] [PubMed]
  234. Mo, F.E.; Muntean, A.G.; Chen, C.C.; Stolz, D.B.; Watkins, S.C.; Lau, L.F. CYR61 (CCN1) is essential for placental development and vascular integrity. Mol. Cell. Biol. 2002, 22, 8709–8720. [Google Scholar] [CrossRef] [PubMed]
  235. Babic, A.M.; Kireeva, M.L.; Kolesnikova, T.V.; Lau, L.F. CYR61, a product of a growth factor-inducible immediate early gene, promotes angiogenesis and tumor growth. Proc. Natl. Acad. Sci. USA 1998, 95, 6355–6360. [Google Scholar] [CrossRef] [PubMed]
  236. Yu, Y.; Gao, Y.; Qin, J.; Kuang, C.Y.; Song, M.B.; Yu, S.Y.; Cui, B.; Chen, J.F.; Huang, L. CCN1 promotes the differentiation of endothelial progenitor cells and reendothelialization in the early phase after vascular injury. Basic Res. Cardiol. 2010, 105, 713–724. [Google Scholar] [CrossRef] [PubMed]
  237. Chintala, H.; Krupska, I.; Yan, L.; Lau, L.; Grant, M.; Chaqour, B. The matricellular protein CCN1 controls retinal angiogenesis by targeting VEGF, Src homology 2 domain phosphatase-1 and Notch signaling. Development 2015, 142, 2364–2374. [Google Scholar] [CrossRef] [PubMed]
  238. Hall-Glenn, F.; de Young, R.A.; Huang, B.L.; van Handel, B.; Hofmann, J.J.; Chen, T.T.; Choi, A.; Ong, J.R.; Benya, P.D.; Mikkola, H.; et al. CCN2/connective tissue growth factor is essential for pericyte adhesion and endothelial basement membrane formation during angiogenesis. PLoS ONE 2012, 7, e30562. [Google Scholar] [CrossRef] [PubMed]
  239. Ivkovic, S.; Yoon, B.S.; Popoff, S.N.; Safadi, F.F.; Libuda, D.E.; Stephenson, R.C.; Daluiski, A.; Lyons, K.M. Connective tissue growth factor coordinates chondrogenesis and angiogenesis during skeletal development. Development 2003, 130, 2779–2791. [Google Scholar] [CrossRef] [PubMed]
  240. Liu, S.C.; Chuang, S.M.; Hsu, C.J.; Tsai, C.H.; Wang, S.W.; Tang, C.H. CTGF increases vascular endothelial growth factor-dependent angiogenesis in human synovial fibroblasts by increasing miR-210 expression. Cell Death Dis. 2014, 5, e1485. [Google Scholar] [CrossRef] [PubMed]
  241. Lin, Z.; Natesan, V.; Shi, H.; Hamik, A.; Kawanami, D.; Hao, C.; Mahabaleshwar, G.H.; Wang, W.; Jin, Z.G.; Atkins, G.B.; et al. A novel role of CCN3 in regulating endothelial inflammation. J. Cell Commun. Signal. 2010, 4, 141–153. [Google Scholar] [CrossRef] [PubMed]
  242. Lin, C.G.; Leu, S.J.; Chen, N.; Tebeau, C.M.; Lin, S.X.; Yeung, C.Y.; Lau, L.F. CCN3 (NOV) is a novel angiogenic regulator of the CCN protein family. J. Biol. Chem. 2003, 278, 24200–24208. [Google Scholar] [CrossRef] [PubMed]
  243. Zhang, C.; van der Voort, D.; Shi, H.; Zhang, R.; Qing, Y.; Hiraoka, S.; Takemoto, M.; Yokote, K.; Moxon, J.V.; Norman, P.; et al. Matricellular protein CCN3 mitigates abdominal aortic aneurysm. J. Clin. Investig. 2016, 126, 1282–1299. [Google Scholar] [CrossRef] [PubMed]
  244. Chen, P.C.; Cheng, H.C.; Wang, J.; Wang, S.W.; Tai, H.C.; Lin, C.W.; Tang, C.H. Prostate cancer-derived CCN3 induces M2 macrophage infiltration and contributes to angiogenesis in prostate cancer microenvironment. Oncotarget 2014, 5, 1595–1608. [Google Scholar] [CrossRef] [PubMed]
  245. Liu, H.; Dong, W.; Lin, Z.; Lu, J.; Wan, H.; Zhou, Z.; Liu, Z. CCN4 regulates vascular smooth muscle cell migration and proliferation. Mol. Cells 2013, 36, 112–118. [Google Scholar] [CrossRef] [PubMed]
  246. Chuang, J.Y.; Chen, P.C.; Tsao, C.W.; Chang, A.C.; Lein, M.Y.; Lin, C.C.; Wang, S.W.; Lin, C.W.; Tang, C.H. WISP-1 a novel angiogenic regulator of the CCN family promotes oral squamous cell carcinoma angiogenesis through VEGF-A expression. Oncotarget 2015, 6, 4239–4252. [Google Scholar] [CrossRef] [PubMed]
  247. Myers, R.B.; Rwayitare, K.; Richey, L.; Lem, J.; Castellot, J.J., Jr. CCN5 Expression in mammals. III. Early embryonic mouse development. J. Cell Commun. Signal. 2012, 6, 217–223. [Google Scholar] [CrossRef] [PubMed]
  248. Butler, G.S.; Connor, A.R.; Sounni, N.E.; Eckhard, U.; Morrison, C.J.; Noel, A.; Overall, C.M. Degradomic and yeast 2-hybrid inactive catalytic domain substrate trapping identifies new membrane-type 1 matrix metalloproteinase (MMP14) substrates: CCN3 (Nov) and CCN5 (WISP2). Matrix Biol. 2016. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of key extracellular matrix (ECM) components involved in angiogenesis. On the left the major classes of ECM proteins that exert a role in angiogenesis are reported. The mechanism of growth factor release from ECM molecules through the action of proteases, activated upon angiogenic stimulus, is highlighted within the baby blue rectangle.
Figure 1. Schematic representation of key extracellular matrix (ECM) components involved in angiogenesis. On the left the major classes of ECM proteins that exert a role in angiogenesis are reported. The mechanism of growth factor release from ECM molecules through the action of proteases, activated upon angiogenic stimulus, is highlighted within the baby blue rectangle.
Ijms 17 01822 g001
Figure 2. Schematic representation of the expression and degradation of MMRN2 during the steps involved in angiogenesis. (1) Following an angiogenic stimulus the basement membrane and MMRN2 are degraded by MMPs; (2) endothelial tip cells which drive sprouting angiogenesis are formed; (3) stalk cell proliferate and vessels’ lumen is formed; (4) ECs secrete MMRN2 stabilizing the vessels; (5) the quiescent state is restored.
Figure 2. Schematic representation of the expression and degradation of MMRN2 during the steps involved in angiogenesis. (1) Following an angiogenic stimulus the basement membrane and MMRN2 are degraded by MMPs; (2) endothelial tip cells which drive sprouting angiogenesis are formed; (3) stalk cell proliferate and vessels’ lumen is formed; (4) ECs secrete MMRN2 stabilizing the vessels; (5) the quiescent state is restored.
Ijms 17 01822 g002
Figure 3. EMILIN2 stimulates angiogenesis via an RTK-dependent cytokine production. Schematic representation of the molecular mechanisms elicited by EMILIN2. The molecule interacts directly with membrane receptors present in both ECs and fibroblast. This leads to the activation of an intracellular signaling cascade that results in the overproduction of angiogenic cytokines that, in turn, boost EC proliferation and migration.
Figure 3. EMILIN2 stimulates angiogenesis via an RTK-dependent cytokine production. Schematic representation of the molecular mechanisms elicited by EMILIN2. The molecule interacts directly with membrane receptors present in both ECs and fibroblast. This leads to the activation of an intracellular signaling cascade that results in the overproduction of angiogenic cytokines that, in turn, boost EC proliferation and migration.
Ijms 17 01822 g003

Share and Cite

MDPI and ACS Style

Mongiat, M.; Andreuzzi, E.; Tarticchio, G.; Paulitti, A. Extracellular Matrix, a Hard Player in Angiogenesis. Int. J. Mol. Sci. 2016, 17, 1822. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17111822

AMA Style

Mongiat M, Andreuzzi E, Tarticchio G, Paulitti A. Extracellular Matrix, a Hard Player in Angiogenesis. International Journal of Molecular Sciences. 2016; 17(11):1822. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17111822

Chicago/Turabian Style

Mongiat, Maurizio, Eva Andreuzzi, Giulia Tarticchio, and Alice Paulitti. 2016. "Extracellular Matrix, a Hard Player in Angiogenesis" International Journal of Molecular Sciences 17, no. 11: 1822. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17111822

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop