Proteoglycan Chemical Diversity Drives Multifunctional Cell

23 hours ago - Biochemistry, Biochemical Analysis & Matrix Pathobiology Research Group, Laboratory of Biochemistry, Department of Chemistry, Universit...
1 downloads 0 Views 18MB Size
Review pubs.acs.org/CR

Cite This: Chem. Rev. XXXX, XXX, XXX−XXX

Proteoglycan Chemical Diversity Drives Multifunctional Cell Regulation and Therapeutics Nikos K. Karamanos,*,†,‡ Zoi Piperigkou,†,‡ Achilleas D. Theocharis,† Hideto Watanabe,§ Marco Franchi,∥ Steṕ hanie Baud,⊥ Steṕ hane Breź illon,● Martin Götte,@ Alberto Passi,∇ Davide Vigetti,∇ Sylvie Ricard-Blum,▲ Ralph D. Sanderson,¶ Thomas Neill,○ and Renato V. Iozzo○

Chem. Rev. Downloaded from pubs.acs.org by UNIV OF FINDLAY on 09/11/18. For personal use only.



Biochemistry, Biochemical Analysis & Matrix Pathobiology Research Group, Laboratory of Biochemistry, Department of Chemistry, University of Patras, Patras 26110, Greece ‡ Foundation for Research and Technology-Hellas (FORTH)/Institute of Chemical Engineering Sciences (ICE-HT), Patras 26110, Greece § Institute for Molecular Science of Medicine, Aichi Medical University, Aichi 480-1195, Japan ∥ Department for Life Quality Studies, University of Bologna, Rimini 47100, Italy ⊥ Université de Reims Champagne-Ardenne, Laboratoire SiRMa, CNRS UMR MEDyC 7369, Faculté de Médecine, 51 rue Cognacq Jay, Reims 51100, France ● Université de Reims Champagne-Ardenne, Laboratoire de Biochimie Médicale et Biologie Moléculaire, CNRS UMR MEDyC 7369, Faculté de Médecine, 51 rue Cognacq Jay, Reims 51100, France @ Department of Gynecology and Obstetrics, Münster University Hospital, Münster 48149, Germany ∇ Department of Medicine and Surgery, University of Insubria, Varese 21100, Italy ▲ University Claude Bernard Lyon 1, CNRS, UMR 5246, Institute of Molecular and Supramolecular Chemistry and Biochemistry, Villeurbanne 69622, France ¶ Department of Pathology, Comprehensive Cancer Center, University of Alabama at Birmingham, Birmingham, Alabama 35294, United States ○ Department of Pathology, Anatomy and Cell Biology, Sidney Kimmel Medical College, Thomas Jefferson University, Philadelphia, Pennsylvania 10107, United States ABSTRACT: The extracellular matrix (ECM) constitutes a highly dynamic threedimensional structural network comprised of macromolecules, such as proteoglycans/ glycosaminoglycans (PGs/GAGs), collagens, laminins, fibronectin, elastin, other glycoproteins and proteinases. In recent years, the field of PGs has expanded rapidly. Due to their high structural complexity and heterogeneity, PGs mediate several homeostatic and pathological processes. PGs consist of a protein core and one or more covalently attached GAG chains, which provide the protein cores with the ability to interact with several proteins. The GAG building blocks of PGs significantly influence the chemical and functional properties of PGs. The primary goal of this comprehensive review is to summarize major achievements and paradigm-shifting discoveries made on the PG/GAG chemistry-biology axis, focusing on structural variability, structure− function relationships, metabolic, molecular, and epigenetic mechanisms underlying their synthesis. Recent insights related to exosome biogenesis, degradation, and cell signaling, their status as diagnostic tools and potential pharmacological targets in diseases as well as current applications in nanotechnology and biotechnology are addressed. Moreover, issues related to docking studies, molecular modeling, GAG/PG interaction networks, and their integration are discussed.

CONTENTS 1. Introduction to the Extracellular Matrix (ECM) Network 2. Proteoglycan Types and Structures 2.1. Proteoglycans as Structural and Biological Players of ECM 3. Glycosaminoglycans: Biosynthesis and Structure−Function Relationships © XXXX American Chemical Society

3.1. GAG Interactions with Proteins and Growth Factors: Biological Functions 3.2. GAGs as Therapeutic Agents and Diagnostic Tools 3.3. Advances in Nano-GAG Technology

B D

I K L

F Received: June 4, 2018

F A

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews 4. Extracellular Proteoglycans 4.1. Versican: Structural Characteristics and in Vivo Functions 4.1.1. Structure and Functional Domains 4.1.2. In Vivo Roles in Development and Disease 4.2. Decorin is a Multifunctional Proteoglycan 4.2.1. Structural Considerations of Decorin 4.2.2. Tumor Suppression via pan-RTK Inhibition 4.2.3. Decorin-Evoked Catabolism: Core of the Proteoglycan Iceberg 4.2.4. Decorin Evokes Endothelial Cell Autophagy in a Peg3-Dependent Manner 4.2.5. Decorin Evokes Tumor Cell Mitophagy in a Mitostatin-Dependent Manner 4.3. Lumican Structure−Function Relationship: Key Findings by Molecular Modeling 4.3.1. Lumican Regulates Tumor Progression in a Tissue-Specific Manner 4.3.2. Lumican Modeling Reveals Specific Interactions with Its Mediators 5. Basement Membrane Proteoglycans 5.1. Endorepellin: Dual Receptor Antagonism, Plethora of Functions 5.2. Endorepellin Suppresses Angiogenesis as a Dual Receptor Antagonist 5.3. Endorepellin Evokes Endothelial Cell Autophagy for Sustained Angiostasis 5.4. Endorepellin Evokes a Pro-Autophagic Gene Signature 6. Cell-Surface Proteoglycans 6.1. Syndecans, Glypicans, and Transmembrane CSPGs: Primer to Cell-Surface PGs 6.2. Cell-Surface PGs as Pleiotropic Integrators of Signaling Processes 6.3. From Ectodomain Shedding to PDZ-Domain Interactions: Molecular Modes of Regulating Cell Behavior 6.4. Cell-Surface PGs as Diagnostic Markers and Therapeutic Targets in Malignant, Inflammatory, and Neurodegenerative Diseases 6.4.1. CSPG4: Prognostic Potential and Target for Immunotherapy in Malignancies 6.4.2. Complex Roles of Glypicans in Diseases and Their Targeting 6.4.3. Syndecans: Potent Regulators of Malignancies, Inflammatory, and Neurodegenerative Diseases 6.4.4. Compounds Interfering with the Functions of Cell-Surface PGs Functions 7. Intracellular Proteoglycans: Functional Roles of Serglycin Serglycin Regulates Immune System Response and Tumorigenesis 8. Regulation of Exosome Biogenesis by Syndecans and Heparanase 8.1. Syndecans Promote Syndecan/Syntenin/ Alix Complex Formation 8.2. Heparanase Trimming of Syndecan HS Chains Enhances Exosome Biogenesis 9. Hyaluronan-Metabolism, Epigenetics, and Functions

Review

9.1. Hyaluronan Metabolism: Overview 9.1.1. Control of Hyaluronan Synthesis 9.1.2. Energy Supply and Hyaluronan Synthesis 9.1.3. Hyaluronan Degradation 9.1.4. Epigenetics of Hyaluronan Metabolism 9.2. Biological Functions of Hyaluronan and Implications in Human Pathologies 9.3. Hyaluronan and Its Biotechnological Applications 10. PG/GAG Interactomes: Topological, Structural, and Functional analysis 10.1. Identification and Characterization of Protein-GAG Interactions 10.1.1. Biophysical Techniques 10.1.2. Computational Studies of Protein− GAG Interactions 10.1.3. GAG Microarrays 10.2. Interaction Networks of Glycosaminoglycans 10.3. Interaction Networks of Proteoglycan Classes 10.4. Integration of Structural, Kinetics, Affinity, and Expression Data in GAG/PG Interactomes 11. Conclusions and Perspectives Author Information Corresponding Author ORCID Notes Biographies Acknowledgments Abbreviations References

N N N O Q Q R S T U V V W X X Y Y Y Z Z AA

AL AM AO AP AP AQ AR AT AT AT AV AV AV AW

AW AW AX AX AX AX AX AY AZ BA

1. INTRODUCTION TO THE EXTRACELLULAR MATRIX (ECM) NETWORK All cells present in tissues and organs are embedded or anchored in specialized three-dimensional (3D) scaffolds called extracellular matrices (ECMs). ECMs are complex networks made up of interconnected macromolecules. They provide specialized niches that affect cellular phenotypes and properties including proliferation, migration, and survival.1,2 Major ECM macromolecules are proteoglycans/glycosaminoglycans (PGs/GAGs), collagens, laminins, fibronectin, elastin, other glycoproteins and proteinases. The composition of macromolecules varies among the multitude of ECMs found in various tissues and organs. The ECM scaffold provides certain biomechanical properties and regulates cell behavior in every tissue type and organ.2,3 The complex ECM types, their dynamic structural and functional multipotency, and structural macromolecules are presented in this section. ECMs are binned in two major classes: the pericellular and interstitial matrices. The pericellular matrices surround specific cell types, creating a layer that is in tight contact with cells or forming a thicker layer that anchors epithelial and endothelial cells called basement membrane.4 The typical structure of basement membranes consists of two networks of laminins and collagen type IV that interact with other embedded molecules within the layers. These components interact with protein networks characterized by collagen type XV and XVIII, PGs such as perlecan and agrin and nidogens.4 Cells attach to

AE

AF AF AF

AG AH AH AH AI AI AK AL B

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Table 1. Extracellular PGs: Schematic Representation of Their Structure, GAG Chain Types, and Modular Domains of the Protein Cores

establish an appropriate functional meshwork upon wound healing and to maintain tissue homeostasis.2 This physiological process is disrupted in pathological situations leading to accumulation of abnormal ECMs that either trigger or facilitate disease progression. The dynamic interplay between modulated ECM and resident cells is a critical process in abnormal wound healing and numerous diseases, including fibrosis, cancer, and atherosclerosis.9−11 For example, tissue organization field theory proposed that cancer is a tissue-based disease facilitated by altered interactions between ECM components and resident cells and adjacent epithelial cells.12 Experimental evidence support a key role of stroma in tumorigenesis.12−16 In recent years, the field of proteoglycans has and continues to rapidly expand. PGs control numerous normal and pathological processes, including ECM supramolecular assembly, morphogenesis, cell signaling, immune response, tissue repair, angiogenesis and cancer progression. PGs at cell surfaces act as integrators of signaling events governing these processes, acting as coreceptors for growth factors, facilitators of chemokine signaling through their G-protein coupled receptors, and modulators of stem cell function.6,17,18 Fundamental cellular behavior such as adhesion, migration, cell signaling, growth, and survival are mediated by the

basement membranes via cell surface integrins to laminin that is close to these cells, and subsequent collagen type IV network connects the overall structure with underlying connective tissue.2,4 Interstitial ECMs are mainly composed of fibrillar collagen types I, II, and III, PGs, the GAG hyaluronan (HA), and many other proteins. They create a meshwork called stroma that contains blood vessels, lymphatics, and nerves that nourishes fibroblasts, immune cells, pericytes, and endothelial cells.1,2 Various cell surface receptors, such as cell surface PGs, integrins, CD44, and discoidin domain receptors (DDRs) receive signals triggered by interactions with ECM components, thus controlling cell functions and fates.3,5 Furthermore, ECM components are capable of interacting with other signaling molecules including growth factors and cytokines creating an ECM reservoir for these bioactive molecules.1,2,6 ECMs are dynamic constructs subject to remodeling under physiological and pathological circumstances. Numerous ECM degrading enzymes including matrix metalloproteinases (MMPs), a disintegrin and metalloproteinases (ADAMs), ADAMs with thrombospondin motifs (ADAMTs), plasminogen activators, hyaluronidases, and heparanase (HPSE) are involved in ECM remodeling.7,8 All resident cells, mostly fibroblasts, are in charge of producing ECM components to reC

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Table 2. Pericellular PGs: Types, Structural Features, GAG Types, and Protein Cores’ Modular Domains

differential expression of PGs on cancer cells and tumor stroma. Post-translational modifications of the PG protein core and secondary interactions with ECM enzymes further influence the biological roles of PG.19−21 PGs consist of a protein core onto which one or more GAG chains are covalently attached. The GAG building blocks of PGs present a unique structural diversity in the saccharide composition, the size and sulfation patterns. It is well-established that the chemical composition of GAGs significantly influence the overall functional properties of PGs. Their size and sulfation pattern, forming a glycocode deciphered by GAG-binding proteins, provide specific regulatory roles for each GAG protein core combination.22,23 In the case of cell surface PGs, the functional diversity can be enhanced by the process of proteolytic ectodomain shedding, which converts membranebound coreceptors into soluble paracrine effectors.20 On the other hand, HA is unique insofar as it lacks sulfate groups and is not covalently bound to a protein core. HA synthesis occurs at cell membrane by enzymes different from those involved in the synthesis of the sulfated GAGs. It is of remarkable importance that HA size regulates inflammation and cell behavior via specific HA receptors including CD44 and HAmediated motility receptor (RHAMM).24 Given that GAGs and PGs interact with many proteins, it is important to build their global interaction networks and tissue-specific interaction subnetworks in order to understand how they work in a concerted fashion in vivo and to decipher the molecular mechanisms underlying their functions. Integrating available structural data permits construction of 3D networks to identify the precise glycocodes regulating specific physiological and pathological processes, to discriminate competitive and

noncompetitive interactions, and to select GAG−protein interactions as therapeutic targets.25,26

2. PROTEOGLYCAN TYPES AND STRUCTURES PGs are expressed by all cells and found in all ECMs. Their GAG constituents are linear polymers consisting of repeating disaccharide units that can be modified with sulfate groups at various positions.6,18,22,27 There are six different types of GAGs that vary in disaccharide unit composition and degree of sulfation. They are assembled via discrete biosynthetic pathways by the coordinated action of a variety of enzymes in the endoplasmic reticulum and Golgi apparatus. The GAG family contains chondroitin sulfate (CS), dermatan sulfate (DS), keratan sulfate (KS), heparan sulfate (HS), heparin (Hep), and HA.22 PGs are extremely heterogeneous macromolecules that are classified as extracellular, pericellular, cellsurface-associated and intracellular ones, according to their localization. They are further classified in subcategories according to their gene homology, structural and functional properties, and modular arrangement. This section provides structural information regarding the PG conformation in terms of protein core and GAG chains. The structure-based concept of PG categories, their GAG chain type, and protein core modular domains are presented in Tables 1−3. Extracellular PGs are found in interstitial ECMs and are classified into two subcategories: hyalectans and small leucinerich PGs (SLRPs).6,17 Hyalectans consist of four PGs, aggrecan, versican, brevican, and neurocan. They possess globular domains at the N- and C-termini by which they bind various ECM components, including HA, forming large aggregates. Hyalectans exhibit significant structural differences D

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Table 3. Types and Structural Characteristics of Cell-Surface Associated (Transmembrane and Anchored) and Intracellular PGs

superfamily, collagens type XV and collagen type XVIII (Table 2).2,6,17 Perlecan and agrin contain an assortment of protein modules and carry up to three HS/CS chains (perlecan) or up to three HS (agrin). Collagen type XVIII possesses up to three HS chains, whereas collagen type XV bears CS chains. Interestingly, C-terminal globular domains of perlecan and collagen type XVIII named endorepellin and endostatin, respectively, are proteolytically cleaved and exert potent proautophagic and antiangiogenic activities.2,6,17 Cell surface PGs encompass two main subfamilies called syndecans and glypicans and other single PGs such as CSPG4, phosphacan, and betaglycan (Table 3). Four syndecans exist (syndecan-1, -2, -3, and -4) and are widely expressed in various cells and tissues with the exception of syndecan-3 that appears to be expressed mainly in neural tissue.6,20 They have an ectodomain, a transmembrane domain, and a short cytoplasmic tail. Up to three HS chains decorate their ectodomain, while syndecan-1 and syndecan-3 can also bear CS chains. The transmembrane and cytoplasmic domains exhibit high homology among the family members. The cytoplasmic domain consists of three distinct regions, two conserved domains, C1 and C2, and a variable domain, V, that possesses several binding motifs including a postsynaptic density 95/ disc-large/zona occludens-1 (PDZ)-binding domain present in C2 region.6,20 Glypicans consist of six members anchored to the cell membrane via a C-terminal glycosylphosphatidylinositol (GPI)-anchor (Table 3). They contain 14 cysteine residues homologous with the cysteine-rich domain of Frizzled proteins. Glypicans are modified with up to three HS chains near the juxtamembrane region and expressed mainly by mesenchymal and epithelial cells.6,19 The single-pass transmembrane PGs, CSPG4, betaglycan, and phosphacan, incorporate a unique set of modular domains into their core proteins. CSPG4 carries up to 3 CS chains and has laminin G-like domains and CS chains attached to its ectodomain. Furthermore, CSPG4 presents with a cytoplasmic PDZ-binding domain. CSPG4 is expressed by melanocytes, stem cells, and glial progenitors.6 Betaglycan contains a PDZbinding domain in its cytoplasmic tail, a zona pellucida

and tissue-specific distribution. Aggrecan (Acan) is mainly found in cartilage but also in perineuronal nets of the brain. It holds approximately 100 potential GAG attachment sites on its core protein and is heavily glycosylated with CS and KS chains. Versican (Vcan) is another member of hyalectans present mainly in interstitial ECMs. Four main Vcan isoforms (V0, V1, V2, and V3) have been found that arise from alternative splicing of exons 7 and 8 encoding the central core protein. Alternatively spliced exons 7 and 8 encode two large subdomains, GAGα and GAGβ, where CS/DS chains are attached. The larger isoform V0 contains both GAGα and GAGβ domains. V1 and V2 isoforms hold GAGβ and GAGα, respectively, whereas V3 lacks both GAG attachment subdomains and is not considered as a true PG. Brevican and neurocan are expressed in the brain holding a protein core with few (one to seven) CS chains attached (Table 1).6,17 SLRPs constitute a family of widely expressed PGs present in most ECMs and tissues. Their relatively low molecular weight and presence of shared structural motifs, the leucinerich repeats (LRRs), are defining characteristics.28 They divide into five classes (I, II, III, IV, and V), according to their evolutionary conservation, gene and protein homology, and chromosomal organization. Classes I (decorin, biglycan, asporin, ECM2, and ECMX), II (fibromodulin, lumican, PRELP, and keratocan), and III (epiphycan, opticin, and osteoglycin) are considered as canonical SLRPs, whereas classes IV (chondroadherin, nyctalopin, and Tsukushi) and V (podocan and podocan-like 1) are noncanonical SLRPs.6,17 They have a small core protein of 36−42 kDa and a central leucine-rich repeats. Class I−III SLRPs exhibit another common structural characteristic, the presence of a long penultimate leucine-rich repeat called the “ear” repeat. Decorin and biglycan are prototype members of SLRPs carrying one or two CS/DS chains, respectively, whereas other SLRPs such as lumican, fibromodulin, and keratocan bear two to four KS chains. SLRPs similarly to other proteins containing small leucine-rich repeats have the tendency to interact with a plethora of other molecules (Table 1).6,17 The subfamily of pericellular PGs encompass two modular PGs: perlecan and agrin, and two members of the collagen E

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Upon matrix remodeling, they are released following degradation of PGs by proteolytic enzymes and HPSE.4,6,37 They exert dual functions acting as pro-angiogenic and antiangiogenic factors. For example, perlecan via its HS chains present at N-terminal domain binds VEGFA and fibroblast growth factors (FGFs) presenting them at their receptors to enhance pro-angiogenic signaling. On the other hand, endorepellin from the C-terminus of perlecan and endostatin from the C-terminus of collagen type XVIII, released upon proteolytic degradation, bind several integrin subtypes and VEGFR2 on the endothelial cell surface and potently inhibit endothelial cell migration and angiogenesis.4,6,37 Cell surface PGs are implicated in cell-cell and cell-matrix interactions affecting cell properties. Syndecans, through HS/ CS chains, bind a plethora of matrix molecules, growth factors, cytokines, and proteolytic enzymes regulating cell adhesion, migration, signaling, and pericellular proteolysis. Syndecans interact with growth factor receptors and integrins through their ectodomain and are linked a cell’s cytoskeleton and signaling molecules through their cytoplasmic tail to promote cell-matrix communication and signaling.20 For example, syndecans form signaling complexes with growth factor receptors and growth factors [for instance the syndecan-1fibroblast growth factor 2 (FGF2)-FGFR1 signaling complex] acting as coreceptors presenting growth factors to their highaffinity receptors more efficiently.38 In addition, syndecan-1 and syndecan-4 form signaling complexes on the cell surface with EGFR and integrins α3β1 and α6β4 to augment EGFRdependent cancer cell migration and survival.39,40 It has been shown that syndecans cooperate with numerous integrins to promote binding to several matrix components and activation of integrin-mediated signaling and cytoskeletal regulation.6,20 Glypicans are involved in growth factors binding, such as Wnts, hedgehog (Hh), and FGFs via HS chains affecting their interaction with cognate receptors and downstream signaling.6,18 Glypican-1 forms a complex between the α3 chain of collagen type V and FGF-2 that is essential for efficient FGF-2 signaling in tumor cells.41 Glypican-3 activates Wnt signaling in hepatocellular carcinoma,42 whereas it inhibits Wnt/β-catenin, phosphatidylinositide 3-kinase (PI3K)/AKT, and mitogenactivated protein kinase (MAPK)/FoxM1 signaling in various malignancies inhibiting their progression.43−46 On the other hand, glypican-3 inhibits Hh signaling by binding Hh, and thereby abrogating binding to Patched.47 Glypicans display context-dependent functions in regulating cell signaling.6 CSPG4 also displays a variety of functions affecting cell signaling, adhesion, migration, invasion, and metastasis.6 CSPG4 cooperates with integrin α2β1 upon collagen type VI binding to activate the PI3K pathway in sarcoma cells.48 Residues, specifically Thr2256 and Thr2314, in the cytoplasmic domain of CSPG4 are phosphorylated by protein kinase Cα (PKCα) and extracellular signal-regulated kinase (ERK) in a integrin β1-dependent manner to regulate tumor cell proliferation and migration.49,50 CSPG4 forms complexes with MMP-2 and membrane type 3 MMP (MT3-MMP) on melanoma cell surface and facilitates MMP-2 activation by MT3-MMP in a CS-dependent manner.51

module, and one HS/CS chain. It is ubiquitously expressed at the cell surfaces and is a coreceptor for the transforming growth factor β (TGF-β) superfamily.6,17 Phosphacan is mainly expressed in the brain. Structurally, it has up to five CS/DS chains, two intracellular catalytic protein tyrosine phosphatase domains D1/2, and an ectodomain that holds an alpha carbonic anhydrase module and a fibronectin type III domain.6,17 At the level of intracellular PGs, serglycin is widely expressed in hematopoietic, endothelial, and smooth muscle cells as well as in fibroblasts. It carries up to eight CS/DS/HS/Hep chains all attached in a narrow part of its protein core characterized by consecutive Ser/Gly repeats.6,29 2.1. Proteoglycans as Structural and Biological Players of ECM

PGs are implicated in the structural organization of ECMs by interacting with multiple matrix components acting as crosslinkers. Selected examples regarding the functional relationship between PGs types and interactions with different molecules such as GAGs, proteins, cell surface receptors, growth factors, and chemokines are presented below. Acan binds HA and link proteins via its G1 domain, forming large supramolecular aggregates creating hydrated gels that provide cartilage with resiliency and viscoelastic properties (Table 1). Acan interacts with simple sugars such as fucose and galactose present on collagen type II or other matrix molecules through its C-type lectin domain on the C-terminal G3 module, thus acting as a molecular bridge that links matrix components.17,30 Similarly, Vcan forms large aggregates with HA but also interacts with a large variety of molecules via its G3 domain during wound healing and inflammation.31,32 Elevated Vcan is present in inflammation and cancer to promote cell adhesion, migration, proliferation, and spreading via binding of its G3 domain to CD44, integrin β1, P-selectin, and toll-like receptors (TLRs).6,31,32 In addition, Vcan activates epidermal growth factor receptor (EGFR) signaling via its EGF-like repeats within the G3 domain promoting cancer cell growth, survival, migration, and invasion.33,34 More insights into Vcan structure and functions are given in section 4.1. SLRPs are an important family of PGs with diverse regulatory roles in ECM assembly and signaling. SLRPs interact with fibrillar collagens, regulating collagen fibrillogenesis by noncovalent binding to an intraperiodic site on the surface of collagen fibrils. The attached GAG chain is critical for collagen fibril alignment and orchestrates proper matrix organization to maintain corneal transparency and tissue biomechanics.6,17 Decorin, biglycan, fibromodulin, and lumican bind a variety of growth factors including TGF-β, platelet derived growth factor (PDGF), vascular endothelial growth factor A (VEGFA), connective tissue growth factor, and cell surface receptors including EGFR, c-Met, IGF-IR (insulin-like growth factor receptor I), vascular endothelial growth factor receptor 2 (VEGFR2), TLRs, via specific LRRs within their protein core, affecting a plethora of biological processes including development, inflammation, autophagy, angiogenesis, and tumorigenesis.6,17,35,36 They differentially affect cell signaling and properties in a context- and molecular-dependent manner acting either as promoters or as suppressors of angiogenesis and tumor development.17,36 Pericellular PGs interact with cytokines and growth factors via their HS/CS chains and sequester them within the ECM.

3. GLYCOSAMINOGLYCANS: BIOSYNTHESIS AND STRUCTURE−FUNCTION RELATIONSHIPS The GAG building blocks of PGs are negatively charged heteropolysaccharides. The saccharide composition of the GAG disaccharide repeating units, the number of units and the F

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 1. Major GAG disaccharides and sulfation sites. GAGs [except HA, (C)] are covalently attached to the proteoglycan core through an oligosaccharide link. (A) CS, DS (GlcA is epimerized to IdoA), (B) HS, and Hep are covalently linked to a serine (Ser) residue via a common tetrasaccharide link consisting of Xyl-Gal-Gal-GlcA. (D) KS I (mainly found in cornea, embryonic lung, and liver) is linked to L-Asn residues via Nglycosidic linkage, KS II (located in bone, cartilage and skin) is attached to L-Ser or L-Thr via O-glycosidic bond, and KS III (found in brain) is linked to L-serine residues through O-glycosidic linkage. CS, chondroitin sulfate; DS, dermatan sulfate; HA, hyaluronan; Hep, heparin; HS, heparan sulfate; GlcA, glucuronic acid; IdoA, iduronic acid; GalNAc, N-acetylgalactosamine; and GlcNAc, N-acetylglucosamine.

sulfation patterns define GAGs classification. In this section, GAG structures are introduced in terms of structure, biosynthesis, and biological functions. The basic structural features of GAGs are schematically presented in Figure 1.

CS/DS/HS and Hep biosynthesis is initiated in the Golgi apparatus by the addition of xylose (Xyl) to a serine hydroxyl group that is invariably followed by a glycine residue on the core protein. Two xylosyltransferases, Xylt1 and Xylt2, catalyze G

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

presented in DS (see below) could be liberated as unsaturated Δ4,5-disaccharides on GlcA/IdoA residues using enzymatic digestions with specific chondroitin lyases (chondroitnases ABC, B, and/or C) and identified following separation techniques, such as HPLC and capillary electrophoresis.54−56 The stereoisomeric variant of CS polysaccharide containing IdoA instead of GlcA has been designated as DS or CSB. The main disaccharide structures of DS are a nonsulfated unit (IdoA-GalNAc), a monosulfated unit at C-4 position of the GalNAc residue [IdoA-GalNAc(4S)], and a disulfated unit at the C-2 position of IdoA and the C-4 position of GalNAc residues [IdoA(2S)-GalNAc(4S)].57 DS is elongated identically as a CS chain but a few more enzymes are involved in its biosynthesis. Dermatan sulfate epimerases 1 and 2 (DSE1 and DSE2) possess chondroitin-GlcA C5-epimerase activity converting D-GlcA to L-iduronic acid (L-IdoA). Therefore, the major difference between CS and DS is that the latter contains a variable percent of IdoA residues instead of GlcA residues within the chain (3-GalNAcβ1−4-IdoAα1-/3-GalNAcβ1−4-GlcAβ1-). DS chains are more flexible, allowing specific interactions with many other molecules due to the different orientation of the carboxyl group of IdoA.58 Additionally, CHST14 preferably transfers sulfate groups to C-4 position of GalNAc residues linked to IdoA in DS and may act immediately after epimerization of GlcA to IdoA. UST also mainly acts toward IdoA residues in DS, sulfating them at C-2 position to accomplish DS biosynthesis.23,52 Similarly to CS, eight different disaccharide units may occur in DS chains providing them with incredible structural diversity. Digestion with chondroitin lyases and performing ion-pair HPLC to identify and elucidate the DS sequence demonstrates sulfation diversity and distribution of the IdoA/GlcA-containing disaccharides within the chain.55,59 HS chains are polymerized by the EXTL1 and EXTL2, after the initial addition of α4-linked GlcNAc to GlcA of the linkage tetrasaccharide by the action of EXTL3, to assemble the HS backbone consisting of 4-GlcNAcα1−4-GlcAβ1-disaccharide units. The complexes of EXTL1/EXTL2 exhibit substantially higher enzymatic activity than EXTL1 or EXTL2 alone. This backbone undergoes extensive modifications by the action of several enzymes. A percentage of GlcNAc residues is modified by a reaction catalyzed by four N-acetyl glucosamine Ndeacetylase/N-sulfotransferases (NDST1−4) that deacetylase GlcNAc and transfer a sulfate group generating N-sulfated glucosamine (GlcNSO3).52,60 Other modifications include the epimerization of GlcA to IdoA by the action of C5-epimerase D-glucuronyl C5-epimerase (GLCE) that differs from DSE1− 2 involved in DS biosynthesis. Furthermore, GlcA and IdoA residues can be substituted with sulfate groups at C-2 position by a reaction catalyzed by HS 2-O-sulfotransferase 1 (HS2ST1). HS 6-O-sulfotransferases (HS6ST1−3) transfer sulfate groups at C-6 position of GlcNSO3 residues. A variety of HS 3-O-sulfotransferases (HS3ST1, 2, 3a, 3b, 4, 5, and 6) are involved in the addition of sulfated groups to the C-3 position of either GlcNSO3 or nonsulfated glucosamine.23,52 This reaction is controlled by the presence of sulfate groups on other positions on the residue and the epimerization of neighboring uronic acid. Plasma membrane endosulfatases can remove sulfate groups from specific sites altering the overall structure. All these modifications establish HS/Hep as the GAG with the highest negative charge (up to 2.7 sulfates per disaccharide). Furthermore, the modifications of HS/Hep chains usually appear in clusters on the chain creating highly

this process using uridine diphosphate (UDP)-Xyl as a donor in vertebrates. Xylosyltransferases could act in a progressive manner since several PGs present clustered GAG attachment sites on their core proteins. Xylosyltransferases exhibit incomplete processing in some cases since several PGs with multiple GAG attachment sites exhibit a different number of GAG chains in different cells. Two Gal and one D-GlcA residues are attached in a stepwise fashion on the Xyl by β4 galactosyltransferase (β4GALT7), β3 galactosyltransferase (β3GALT6), and β3 glucuronosyltransferase (β3GAT7) to complete the GAG linkage region: GlcAβ1-3Galβ1-3Galβ14Xylβ1-Ser. This tetrasaccharide is further modified by phosphorylation catalyzed by GAG xylosylkinase FAM20B at the C-2 position of Xyl and sulfation of the Gal residues as seen in CS.52 This intermediate tetrasaccharide lies at the bifurcation of CS/DS and HS/Hep biosynthetic pathways. The addition of the next hexosamine residue commits the intermediate to either CS/DS or HS/Hep. A β4-linked Nacetyl-D-galactosamine (GalNAc) is attached to GlcA by the reaction catalyzed by the enzymes CS N-acetylgalactosaminyltransferases 1/2 to initiate the polymerization of CS/DS. On the other hand, an α4-linked N-acetylglucosamine (GlcNAc) is bound to GlcA by the action of exostosin-like glycosyltransferase 3 (EXTL3) for HS/Hep chain polymerization.52 CS is a GAG composed of repeating disaccharide units (-4D-GlcA-β1−3GalNAcβ1-)n and modified with various sulfation patterns on the monosaccharide residues.53 The average molecular weight of CS is 10−100 kDa, consisting of 40−400 monosaccharide residues as a mature molecule. CS chains are elongated by the polymerization of repeating disaccharides of −3-GalNAcβ1−4-D-GlcAβ1- that can be modified with sulfate groups at the C-4 and C-6 positions of GalNAc and in some cases at C-2 of GlcA. The bifunctional enzymes that catalyze the polymerization are chondroitin synthases 1−3 (CHSY1− 3) that possess both β-1,3-GlcA and β-1,4-GalNAc transferase activity using UDP-GlcA and UDP-GalNAc as donors. CHSY2 is also called as chondroitin polymerizing factor (CHPF), and although it lacks an independent activity, it is required for CS biosynthesis as it collaborates with other enzymes promoting polymerization.52 Numerous sulfotransferases are involved in the modification of elongating GAG chains with sulfate groups using 3-phospho-5-adenylyl sulfate (PAPS) as the sulfonate donor. Carbohydrate sulfotransferases 3 and 7 (CHST3) and (CHST7) catalyze the transfer of sulfate group to C-6 position of GalNAc, whereas CHST11−13 catalyze the sulfation of GalNAc at C-4 position in CS. CHST15 transfers sulfate group to the C-6 position of GalNAc sulfated at C-4 to form a disulfated GalNAc(4,6S). Finally, uronyl 2-sulfotransferase (UST) has a weak enzymatic activity toward GlcA catalyzing its sulfation to C-2 position.23,52 All these modifications contribute to CS structure diversity, and a combination of eight differently modified disaccharide units may coexist in a CS chain. The typical disaccharide units of CS (i.e., 0, A, C, D, E, and T) correspond, respectively, to the following: nonsulfated unit (GlcA-GalNAc), monosulfated unit at the C-4 position of the GalNAc residue [GlcA-GalNAc(4S)], monosulfated unit at the C-6 position of the GalNAc residue [GlcA-GalNAc(6S)], disulfated unit at the C-2 position of GlcA and the C-6 position of GalNAc residues [GlcA(2S)GalNAc(6S)], disulfated unit at the C-4 and C-6 positions of the GalNAc residue [GlcA-GalNAc(4S,S)], and trisulfated disaccharide unit [GlcA(2S)-GalNAc(4S,S)]. The differently sulfated CS-derived disaccharides as well as those mainly H

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

sulfated clusters with enhanced biological activity. Although both HS and Hep share the same backbone and modifications, they differ in sulfation and epimerization. The differentially sulfated disaccharides could be liberated from the intact Hep/ HS chains using Hep lyases I, II, and III alone or in combination and identified by HPLC and capillary electrophoresis according to their sulfation patterns.61−63 HS exhibits a lower degree of sulfation and epimerization possessing approximately one sulfate group per disaccharide unit and mainly contains GlcA within the chain backbone.23,60 Hep is a more negatively charged and structurally complex relative of HS that is found attached to serglycin core protein synthesized in connective tissue mast cells.29 KS consists of repeating disaccharides of 4-GlcNAcβ1−3Galβ1- polymerized by the catalytic activity of β1−3-Nacetylglucosaminyltransferase-7 (B3GNT7) and β1−4-galactosyltransferase-4 glucosaminyltransferase-7 (B4GALT7).64,65 Disaccharides can be sulfated at C-6 positions of both Gal and GlcNAc. CHST5 and CHST6 catalyze the transfer of sulfate group to C-6 position of GlcNAc, whereas CHST1 catalyzes the sulfation of Gal at C-6 position.23,64 The linkage region of KS on certain PGs differs in a tissue-dependent and PG-dependent manner. KS-I is polymerized on a complex-type N-linked branched oligosaccharide linked to a protein core via an asparagine residue in cornea and cartilage PGs. In cartilage, KS-II is found on an O-glycan mucin core-2 structure linked to a serine or threonine residue via a GalNAc exclusively in Acan. KS-III found in the brain is linked to a serine residue through a mannose residue.52,64 HA is the only GAG type that is synthesized on the inner face of plasma membrane by the action of three isoenzymes called HA synthases (HASs, HAS1−3) that are localized as complexes within the plasma membrane. The precursors UDPN-GlcNAc and UDP-GlcA are polymerized by HASs, and the newly formed HA consisting of (β1−3-N-GlcNAc-β1−4-DGlcA) disaccharides is secreted in the ECM through the HASs complexes.66 HASs synthesize large HA polymers with various sizes. HAS1 and HAS2 produce HA of higher molecular size (2 × 106 Da), whereas HAS3 synthesizes HA of lower size (2 × 105 Da). Notably HA is not modified with sulfate groups as other GAG types, and its biosynthesis is regulated by several modifications occurring in HASs such as N-glycosylation, OGlcNAcylation, and ubiquitination.24

observed that GAG chains may interact with more than one ligand (e.g., Hep with FGF), since they exist in a helical conformation, which means that only GAG residues facing the protein interact with AA residues, whereas residues on the other side of the helix are able to interact with another protein.67 GAG-binding proteins include ECM proteins (e.g., collagens), enzymes (e.g., proteases), and enzyme inhibitors (e.g., antithrombin III), chemokines (e.g., interleukins), growth factors (e.g., FGF), and membrane receptors (e.g., FGFR, integrins), lipid-binding proteins (e.g., apolipoproteins), pathogen surface proteins (e.g., malaria circumsporozoite protein), and viral envelope proteins (e.g., Zika virus and HIV).68 The majority of GAG-binding proteins have been identified by affinity-based approaches, including, affinity chromatography, coprecipitation, fluorescence titration, and surface plasmon resonance, whereas data regarding structural information has been obtained by X-ray crystallography and solution-state nuclear magnetic resonance (NMR).69−71 To a large extent, the majority of GAG-binding proteins interact with Hep or HS, compared to other GAGs, due to the greater sequence heterogeneity and more variable sulfation patterns. The best-studied model of GAG−protein interaction is the binding of antithrombin III to Hep and HS, which has great pharmacological relevance due to Hep’s application as an anticoagulant. Binding of antithrombin III to Hep results in strong conformational changes and enhances the inhibiting rates of thrombin and coagulation factor Xa. Henceforth, Hep chain acting as a catalyst, reinforces the contact of thrombin and antithrombin, inferring that both have GAG-binding sites. However, only about one-third of Hep chains bind with high affinity to antithrombin III, that contribute to this activation/ inactivation cycle, which means the binding sites correspond to a very small segment of the GAG chains, a concept that can be extrapolated to all GAG-binding proteins.69 X-ray crystallography data revealed that antithrombin A (41PEATNRRVW49) and D (124AKLNCRLYRKANKSSKLVSANR145) helices, which contain critical Arg and Lys residues at the interface, contact the docking site of Hep pentasaccharide. Binding to the pentasaccharide is sufficient for antithrombin activation. However, to prevent ternary complex formation (antithrombin/thrombin/Hep), a larger oligosaccharide of 18 residues is needed.72 The role of Hep/HS in cell functions, such as adhesion and migration, is critical. Brunetti and colleagues73 evaluated the effect of a tetra-branched peptide (NT4) on cancer cell adhesion and migration in several in vitro models. They discovered that NT4 binds to HS and inhibits cell adhesion and migration without any evident modification of cancer cell morphology as observed by scanning electron microscopy (SEM). NT4 did not affect the ability of cancer cells to produce cytoplasmic protrusions, suggesting a determinant role of sulfated GAGs in the control of cancer cell directional migration. As mentioned above, HS is the predominant GAG of HSPGs (syndecans 1−4). Chikama and colleagues provided compelling data that syndecan-1 can prevent calcium oxalate monohydrate crystal attachment on renal tubular epithelial cells surface as observed by SEM images.74 Detailed information regarding the Hep/HS interaction network will be addressed in section 10. It is well-established that distinct GAG interactions orchestrate chemokine-dependent neutrophil recruitment in response to microbial infection. Responsive for neutrophil trafficking is the conserved N-terminal Glu-Leu-Arg motif in

3.1. GAG Interactions with Proteins and Growth Factors: Biological Functions

GAG chains represent one of the most investigated ECM molecules that regulate various signaling pathways, normal and pathological processes. Insights into the various types of GAG interactions with growth factors and proteins are given below. Interactions of GAGs with many different classes of proteins, mostly through electrostatic interactions between negatively charged carboxylate and sulfate groups (GAG-binding sites) and positively charged protein amino acids (AA), may have profound effects on physiological processes like development, cell growth, hemostasis, and lipid transport/absorption. The predicted binding sites in proteins contain basic AA (Arg and Lys) that interact, through the positively charged residues in the α-helices and β-strands, with the negatively charged hexuronic acids and sulfate groups of GAGs. This may potentially regulate enzymatic activity, protein oligomerization, and protection against proteolytic degradation as well as the binding of ligands to their receptors. Moreover, it has been I

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

of disulfated units. Sequential reactions with UST and GalNAc4S-6ST generated further highly sulfated CS containing a mixed structure of disulfated units. Sequential reactions with GalNAc4S-6ST and UST generated a novel CS molecule containing 29% trisulfated disaccharide units. The CS library enables evaluation of affinity and specificity of CS-binding molecules and determination of the epitope of anti-CS antibodies. CS polysaccharides, oligosaccharides, and chemoenzymatically synthesized CS (sCSs) are attached with hexamethylene diamine (HMDA) at the reducing end, and biotinylated. They are immobilized onto a streptavidin (SA)chip or onto SA-coated ELISA plates, enabling surface plasmon resonance (SPR) analysis or ELISA for CS-binding.79 When MK and PT were used for SPR analysis, MK bound CS-D and CS-E polysaccharides and sCSAD at a low dissociation constant (Kd). PT binds CS-D and CS-E polysaccharides and sCS-E at low Kd. Epitopes of anti-CS antibodies had been determined by inhibition ELISA, and there had been only a limited number of results showing direct interactions of the antibodies with different CS species. The study using ELISA with synthesized CS species has revealed that the anti-CS antibodies interact more strongly with complex CS structures, such as sCSDE, sCSAD, and sCtriS, compared to uniform sCS species, such as chondroitin (sCH), sCSA, sCSC, and sCSE.79 It is well-established that the molecular basis of GAG− protein interactions is dominated by charges. In the case of HA, the nonsulfated GAG, due to its enormous size, an individual HA can bind simultaneously a large number of proteins (∼100) and to multiple cell surface receptors (∼10), via different types of molecular interactions (ionic, hydrogen bonds, Van der Waals, and aromatic ring stacking).82 Therefore, considerable diversity exists in the interactions of HA-binding proteins (hyaladherins) with HA, which depends on the particular size of HA-binding sites in several proteins, including among others TNF-stimulated gene 6 protein (TSG6), lymphatic vessel endothelial HA receptor 1 (LYVE-1), CD44, and Acan. Co-crystal structures have been determined for HA, the principal ligand of the multifunctional cell receptor and a part-time PG, CD44, and HA-binding domains in the receptor. Moreover, molecular modeling revealed junctions between HA chains and TSG-6.83 In the case of LYVE-1, it only binds HA if the necessary pentasaccharide is wellorganized, for instance, cross-linked with TSG-6.84 Covalently cross-linked HA complexes result from antiparallel arrangement of HA-binding sites mediated by TSG-6 through the transferase reaction. The noncovalently cross-linked HA chains, formed at inflammation sites via pentraxin-3 binding, explain the hydrodynamic properties of the polysaccharide.85 In order to assess the influence of extracellular HA on cell− matrix interactions, human skin fibroblasts were cultured in a 3D collagen gel or in monolayer, in the presence or absence of HMW-HA. Fibroblasts retracted the gel in a time-dependent manner; however, the retraction was not affected by HA concentration, and a low adhesion capacity to matrix HA, either soluble or immobilized, was reported. Therefore, it has been suggested that the ability of dermal fibroblasts to bind collagen is not affected by the presence of HMW-HA. On the other hand, HA seems to have great influence on fibroblast shape and orientation. In fact, transmission electron microscope (TEM) analysis demonstrated that the exogenous HA increases actin cytoskeleton filaments underneath the plasma membrane, suggesting that CD44-mediated signal transduction by HA mainly affects cell fibroblast locomotion and

human chemokines (CXCL1, CXCL2, CXCL3, CXCL5, CXCL6, CXCL7, and CXCL8). Neutrophil-activating chemokines (NACs) bind HS in endothelium, epithelium, and ECMs and exert their functions via CXCR2 signaling on neutrophils. The molecular mechanism by which chemokine monomers/ dimers interact with Hep oligosaccharides and CXCR2 has been deciphered by NMR spectroscopy. NMR studies indicated that basic residues in chemokines structures (Lys, Arg, and His) mediate binding affinity with the highly acidic GAGs.75 CS and the related GAG, DS, are physiologically relevant binding partners, since they are the prevalent GAGs in many tissues. It has been believed that, similarly to HS, CS can bind various molecules mediating its biological functions. Indeed, several molecules bind CS, although its binding affinity is often lower than HS. Especially, as CSPGs are enriched in brain as well as in cartilage, several molecules in nervous system have been investigated for CS interactions.76 For instance, growth factors and chemokines, such as FGF-2, brain-derived neurotrophic factor (BDNF), and interleukin-10 (IL-10), bind to CS-A. FGF-16, FGF-18, and HB-EGF bind to CS-E with a comparative affinity to Hep, in a CS-structure-specific manner rather than via nonspecific charge interactions.77,78 In addition, midkine (MK) and pleiotrophin (PT), two classic Hep-binding proteins are well-known to bind CS. The affinity and specificity of their binding to CS are discussed in this section.79 Three transmembrane receptor protein tyrosine phosphatases (PTPs) have been identified as functional receptors of CS/DS. Of them, PTPσ binds to CS/DSPGs and inhibits axonal growth by inhibiting its oligomerization. Interestingly, this receptor binds HS and facilitates growth by inducing the receptor oligomerization. A recent study revealed that CS-E of ∼10 kDa binds to PTPσ with the highest affinity.79 Furthermore, the receptor for advanced glycation end products (RAGE) located at the cell surface of Lewis lung carcinoma (LLC) cells is the predominant receptor for CS/DS chains containing E units. RAGE recognizes specific sulfation patterns or sequential arrangements, as it strongly interacts with CS-E, via its basic AA region. RAGE also moderately binds DS and CS-D, whereas CS-A and CS-C exhibited moderate binding affinity in vitro.80 As mentioned above, NACs also bind CS and DS, whereas this mechanism is less understood, as CS/DS differ from Hep and contain less sulfation. A recent NMR study indicated that chemokine (C− C motif) ligand 5 (CCL5) binds CS oligosaccharides through a positively charged and fully exposed motif, KKWVR.81 At present, commercially available CS species are a mixture of different CS disaccharide units, although CS-A from whale cartilage, CS-C from shark cartilage, CS-D from shark fin cartilage, CS-E from squid cartilage, and DS from pig skin are rich in the corresponding unit, respectively. Provided that CS chains exert function with ∼20-mer stretches like HS, studies using the aforementioned CS polysaccharides may not clarify the exact CS structure required for function. Developing a CS library, containing various CS species with defined lengths and defined sulfate compositions, has been generated by chemoenzymatic synthesis.6 Reactions with chondroitin polymerase generated nonsulfated chondroitin, and those with chondroitin-4-sulfotransferase 1 (C4ST-1) and chondroitin-6-sulfotransferase 1 (C6ST-1) generated uniformly sulfated CS containing >95% 4S and 6S units, respectively. N-Acetylgalactosamine 4-sulfate 6-sulfotransferase (GalNAc4S-6ST) and UST generated highly sulfated CS containing mixed structures J

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

3.2. GAGs as Therapeutic Agents and Diagnostic Tools

orientation, as confirmed by the fusiform shape of these cells in the presence of immobilized HA.86 It is of interest that HMW-HA molecules, at high concentration, tend to aggregate and form long parallel strands to which dermal fibroblasts are not attached, but they are connected and migrate in between them by assuming a very long and thin shape.87 In order to analyze cellular interactions with HA, Dickinson and Gerecht developed functionalized surfaces that enable a reliable study and high-resolution visualization of these interactions. They showed that both MDA-MB-231 breast and LS174t colon cancer cells preferentially adhere to HA-functionalized sites through CD44 and develop adhesive protrusions and spreading by SEM.88 Furthermore, HA seems to counteract the effect of a cytokine involved in cartilage degradation processes, thus preserving the normal shape of human osteoarthritic chondrocytes.89 The authors tested the effects of interleukin-1 β (IL-1β) and HA (500−730 kDa) in the presence and absence of cyclic hydrostatic pressure to investigate the morphology of human osteoarthritic chondrocytes and production of PGs and nitrite ions (NO). Cells cultured only in IL-1β presence showed reduced production of PGs and increase concentration of nitrites; however, HA has a protective role against the negative effects of IL-1β and restored cell structure, with a significant increase in PGs and a decrease in nitrite levels with or without the pressurization. The human osteoarthritic chondrocytes cultivated in the presence of IL-1β and HA, when observed by TEM and SEM, showed normal spherical shapes and the reestablishment of structures related to cell anabolism, such as granules and fibrils. Chondrocytes cultured in monolayered culture lost their typical round morphology and showed a more fibroblast-like shape with increased collagen type I and a decrease of collagen type II and Acan production.90 However, the development of a novel 3D scaffolds of chemically unmodified HA with minimum cross-linkage to promote cartilage repair, demonstrated that spherical-shaped cells form aggregates, have slow proliferation rates, and decreased DNA synthesis. These cells expressed transcripts encoding collagen type II and Acan and produced sulfated GAGs. These data suggest that HA favors normal chondrocytes grown not only for the chemical presence but also for the physical environment residing in a 3D scaffold. Moreover, Turner and colleagues observed by SEM that endothelial cells adhered to the individual fibers of both unpressed and pressed fibrous, nonwoven HA-based scaffolds (Hyaff-11) for tissue repair.91 Pasquinelli and colleagues investigated the similar scaffolds and demonstrated that native HA favors adhesion, migration, and proliferation of rat mesenchymal cells as well as synthesis and delivery of autologous ECM components.92 Another interesting example of GAG−protein interactions includes lumican, a SLRP mainly containing KS, that attenuates proliferation, migration, and invasion of breast cancer cells. Lumican treatment of MDA-MB-231 breast cancer cells and shERβ MDA-MB-231, where estrogen receptor beta (ERβ) has been stably suppressed, had cell− cell junctions and a modified cell shape; in contrast, cancer cells appeared flattened and underwent mesenchymal-toepithelial transition (MET).93 Detailed information regarding the role of the lumican structure−function relationship will be addressed in section 4.3.

GAGs are among the critical ECM mediators that affect cell functions, while their structural motifs account for their ability to directly bind proteins, growth factors cell receptors, and chemokines. The tremendous variations on GAGs structures gain significance as biomarkers and pharmacological targets for disease treatment and fruitful biomedical applications. The comprehension of the GAG structure−function relationship and their numerous roles in biological processes can be accomplished with the development of molecular tools that selectively interfere with GAG structures. Therefore, GAG targeting in biological systems serves as an attractive objective for modern therapeutics and translational medicine. Because of its ability to bind antithrombin, Hep isolated from porcine and bovine intestinal mucosa is the most widely used GAG, applied as an anticoagulant and antithrombotic agent. However, the therapeutic use of unfractionated Hep is limited due to its severe hemorrhagic effects and poor bioavailability. Overcoming these side effects has been achieved by the development of several synthetic and semisynthetic Hep analogs which have been developed, mostly through the degradation of unfractionated Hep. Such analogs are Hep-mimicking compounds, low-molecular-weight Hep, and ultralow-molecular-weight Hep, which have been demonstrated to possesses “natural” anticancer properties in addition to their function as an anticoagulant, such as fondaparinux.94−96 Non-anticoagulant LMW Hep, such as dalteparin, are currently in clinical trials for patients with lung carcinoma, refractory prostate cancer, and pancreatic cancer.97 Besides classical examples, such as fondaparinux, immunomodulatory drugs such as synthetic disulfated IdoA harbor considerable therapeutic potential, as they target GAG-ligand interactions in a specific manner.98 HS-based therapy involves the modification of HS-degrading enzymes activity. Among therapeutic strategies to inhibit HPSE, PI-88, a phosphomannopentaose sulfate, a highly sulfated oligosaccharide mixture, has entered clinical trials II/ III for metastatic melanoma. PI-88 blocks tumor growth and angiogenesis by inhibiting the interactions of FGFR1, FGFR2, and VEGFR2 with their ligands and HS chains.99 Roneparstat (SST0001) is a non-anticoagulant, glycol-split Hep with antiHPSE activity. This modified Hep is a promising tool to overcome abnormal angiogenesis and poor drug delivery and when combined with other conventional drugs (e.g., dexamethasone), which resulted in synergistic antiangiogenetic effects and tumor growth inhibition in vivo.100 Moreover, when evaluated in phase I clinical trials (NCT01764880) for the treatment of advanced multiple myeloma, no dose limiting side effects were identified.101 CS has also been exploited as a pharmacological agent. Due to its critical role in cartilage and other connective tissues, CS represents an alternative therapeutic in osteoporosis and osteoarthritis. In a mouse breast cancer model, tumor growth was abolished by carbodiimide-modified CS chains. 102 Fucosylated CS and oversulfated CS chains possess inhibitory effects on P-selectin-mediated processes.96 Novel, cisplatinloaded PEG-coated liposomes can be internalized into cancer cells when bound to CS, improving drug delivery.103 DS is overexpressed in tumor microenvironment. Therefore, on this basis, a useful molecular tool against cancer progression may be the specific phage display-derived antibodies LKN1 and GD3A12, which selectively recognize IdoA-GalNAc4S disaccharide units in DS polysaccharides.104 Furthermore, recent K

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

plantation.127 Serum concentrations of HA in patients with rheumatoid arthritis and hand osteoarthritis were significantly higher than in healthy donors, suggesting the predictive relevance of HA as a diagnostic tool for monitoring autoimmune pathologies and disease progression.128 Mounting evidence has focused attention on the manipulation of GAG biosynthetic enzymes, by the utilization of synthetic xylosides. GAG composition can be manipulated in several pathologies via synthetic xylosides comprised of a xylose and an aglycone group that competes with GAG biosynthetic enzymes to bind GAGs. This results in reduced PG-bound GAGs and increased xyloside-bound GAGs, mostly CS and HS, thereby modulating their biological functions.129 Since they are small molecules, xylosides can be easily excreted and exhibit oral bioavailability with high diffusivity. They have an advantage over synthetic- or animal-derived GAGs for potential therapeutic applications, as they derive from the natural biosynthetic apparatus, thus being likely nonimmunogenic. An interesting example evaluated in phase II clinical trials refers to a bioavailable β-D-thioxyloside, odiparcil, which significantly downregulated venous thromboembolism in patients with hip or knee replacement.130,131 Moreover, naphthyl-xyloside-primed CS exhibited in vitro cytotoxicity by reducing breast cancer cell growth.132 β-D-Xylosides reduce glioma cell migration and exosome uptake, while acting as inhibitors of HS/CSPGs biosynthesis and abolishing tumorassociated angiogenesis.133,134 Xylosides have significant applications in regenerative medicine; for instance, in spinal cord injury models, where axonal growth improved following the inhibition of CSPGs expression with β-D-xylosides.135 Concluding, further structure−activity elucidation of xylosides is required to achieve the desirable biological effects for the development of novel GAG-based therapeutic applications. Given that GAGs and PGs interact with many proteins (∼450 for HS/Hep), it is important to build their global interaction networks and tissue-specific interaction subnetworks in order to understand how they work in a concerted fashion in vivo and to decipher the molecular mechanisms underlying GAG/PG functions. Integrating structural data in these interactomes (e.g., protein binding sites and GAG features involved in protein interactions such as the number and type of sulfate groups) is useful to build 3D networks, to identify the glycocodes regulating specific physiological and pathological processes, to discriminate competitive and noncompetitive interactions, and to select GAG−protein interactions as therapeutic targets (see section 10).

advances in nanotechnology allows the selective targeting of cancer cells that can be achieved with biodegradable polymers incorporating CS, chitosan polymers cross-linked with CS, and polyethylenimine nanostructures encapsulating CS.22 The measurement of the selective accumulation of total GAGs has been established for the diagnosis of mucopolysaccharidoses. The concentration of urinary GAGs (especially HS and DS) in patients with different clinical background type of mucopolysaccharidosis discriminates the attenuated and severe type of mucopolysaccharidoses.105 Moreover, asporin, a SLRP, is highly expressed in prostate cancer cells and localized in the reactive stroma, and thus may serve as a prognostic biomarker for aggressive cancer subtypes.106 HA is a key mediator of the epithelial-to-mesenchymal transition, a process during which cells establish a migratory phenotype that is critical for the initiation of cancer metastasis. HA expression is inhibited by 4-methylumbelliferone (4-MU), a coumarin derivative with low cytotoxicity that has been approved as a dietary supplement. Moreover, its tumorsuppressive functions in breast cancer cells with different metastatic potential have been demonstrated,107 as well as in the treatment of prostate cancer in preclinical studies.108 The increased HA deposition in tumor sites triggers signaling cascades, thus promoting the interaction between CD44 and other cell receptors. On this basis, the manipulation of HACD44 interactions in tumor sites is postulated to regulate metastatic behavior.109 HA-CD44 interactions can be inhibited either by small HA oligosaccharides, by an antibody against HA-binding site of CD44, or by CD44 siRNAs, which inhibit the post-transcriptional expression of CD44.22 Given that CD44 internalizes HA, several HA-conjugated drugs and HA conjugated nanocarriers have been developed as drug delivery vehicles. The carboxylate group on GlcA, the hydroxyl group of GlcNAc, and the reducing end, are potentially capable of conjugating a drug, which is internalized via CD44-HA-drug interactions and drug release after enzymatic hydrolysis in the cytoplasm.110,111 For instance, the antimitotic chemotherapeutic agent, paclitaxel, when conjugated to HA, showed targeted cytotoxicity, increased cellular uptake and improved solubility.112 It has been demonstrated in preclinical studies that the cytotoxic properties of an anticancer drug can be improved after chemical conjugation to HA. HA-targeted nanocarriers, utilizing liposomes or nanoparticles, have been applied for the delivery of several anticancer agents, such as doxorubicin, epirubicin, mitomycin c, and paclitaxel, to CD44 overexpressing cancer cells with improved efficiency.113−116 The diagnostic relevance of HA is also of great importance. Regarding HA in pleural inflammation effusion, assaying HA levels in serum and plasma predicts early diagnoses of complicated parapneumonic pleural effusions that characterize infectious pleuritis.117 Increased HA in the pleural fluid is a significant prognostic biomarker for malignant mesothelioma and poor prognosis.118−122 Elevated HA is detectable in pancreatic tumors with poor survival, revealing its prognostic impact for pancreatic ductal adenocarcinoma as well.123 HA has also been utilized monitoring diabetic foot ulceration, a significant complication in patients suffering from diabetes mellitus, by increasing the wound healing process.124 HA was identified as a reliable prognostic serum marker of cirrhosis and fibrosis from chronic hepatitis C.125,126 Furthermore, elevated HA serum levels have been identified as marker for endothelial cell injury in patients with veno-occlusive disease, following bone marrow trans-

3.3. Advances in Nano-GAG Technology

The dynamic interplay among ECM macromolecules, such as PGs and GAGs, generates complex structural networks with other ECM components that critically modulates the pathophysiology of several biological processes. The biological, mechanical, and chemical properties of PGs make them crucial constituents of tissue engineering scaffolds.6,136 Advances in understanding the dynamic interactions among matrix components have led to the design and development of a great variety of synthetic scaffolds comprised of synthetic and semisynthetic mimetics. These may serve as chemical and functional therapeutic replacements of natural PGs in tissue engineering applications, anticancer agents, site-specific drug, and gene delivery. These novel formulations are known as neoproteoglycans (neoPGs), and despite some limitations in their functions, such as limited signaling ability and L

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

growth factors, as well as improved cell proliferation and migration in vivo.149 Hep-conjugates as drugs with nanostructure have already been tested as favorable candidates for drug delivery. Hep-doxorubicin is referred as the combination of dendrimers with Hep characterized by proved clinical success, with pharmacokinetic advantages. This dendronized Hepdoxorubicin pH-responsive drug delivery vehicle has important antitumor efficacy in vivo and induces apoptosis with minimal side effects.150,151 HS-chitosan nanoparticles immobilized to the nanofibers of bovine jugular vein scaffolds by ethylcarbodiimide hydrochloride/hydroxysulfosuccinimide modification support cell proliferation and chondrocyte growth in a model of articular cartilage. It stimulates endothelial cell proliferation in vitro, and vascularization, ECM biosynthesis, and fibroblast infiltration in vivo, providing a novel system for accelerated regeneration of tissue-engineering scaffolds.152 Silver-HA nanocomposites obtained by chemical reduction of silver nitrate by hyaluronic fibers exhibited strong antimicrobial effects against Gram positive Staphylococcus aureus153,154 and is used for in vivo imaging after radiolabeling with 99mTc.155 The green synthesis of CS-stabilized silver nanoparticles succeeded from silver nitrate and D(+)-glucose as the reducing agent of Ag+ ions under alkaline conditions. CS silver nanoparticles exhibited outstanding catalytic activity, biocompatibility, low cytotoxicity, and strong antimicrobial functions against Acinetobacter baumannii, Escherichia coli, Pseudomonas aeruginosa, and Staphylococcus aureus.156 CSderived amphiphilic nanoparticles are promising anticancer vehicles, due to their narrow size distribution, high drug loading efficiency, and stabilization by π−π stacking interactions with aromatic antineoplastic agents. These theranostic CS nanoparticles resulted from the conjugation of fluorescein with CS using carbodiimide chemistry and the amine residue of doxorubicin that favors electrostatic binding with the carboxylate/sulfate moiety of CS.157 Multifunctional selfassembled supramolecular CS nanoparticles resulting from the same experimental procedure demonstrated increased cellular uptake and dose-dependent cytotoxicity in different human cancer lines, while triggering complement activation (C3a and sC5b-9), suggesting their novel roles as nanocarriers for cancer therapeutics.158 The design of CS-conjugated nanofibers into poly(vinyl alcohol) scaffolds for engineering of articular cartilage tissue, increased cartilaginous tissue formation, through collagen type II secretion and induced chondrogenesis.159 The design of a multiphase osteochondral scaffold also applied in bone and cartilage regeneration includes the formation of CS-collagen type I and II-calcium phosphate conjugates.160,161 A blood-compatible conjugate of disulfide-cross-linked linear polyethylenimine (used as an efficient gene carrier with low toxicity) and DS exhibited significant cytotoxicity against several cancer cell lines, thus suggesting its potential utility for cancer therapy.162 Another interesting example for tissue engineering applications includes the formation of alginate/chitosan/DS polyelectrolyte microspheres in acidic conditions by spray drying, induced the sustained release of DS, stimulating cell proliferation in vitro.163

biocompatibility, they serve as promising tools in replacing the natural activity of PGs through cell and protein binding interactions. The main types of neoPGs include protein−GAG conjugates, polymer-GAG complexes, and nano-GAG composites.137,138 Nanostructures benefit from the rest, since they can easily penetrate cells and tissue barriers and pass through capillary vessels, with prolonged half-lives in the bloodstream.139 Nano-GAG composites are nanoscale structures with attached GAG chains, where a nanomaterial serves as a substitute for the core proteins. In this section, a structurebased view regarding the types of nanoformulations that encapsulate GAGs, the advantages of their utilization, as well as some examples of their applications in tissue engineering and therapeutics are discussed. The unique physicochemical characteristics of nanomaterials, which make them attractive for numerous applications, include the enormous mass-to-surface ratio, increased solubility and surface reactivity, aggregation, and encapsulation tendency. These novel properties of nanomaterials result in drastically different chemical and biological properties compared to the same material in bulk form.140 The improved surface versatility enables conjugation of a broad range of biomolecules and renders nanoparticle-based drug delivery systems. They are emerging as very promising tools for cancer therapeutic approaches, as nanoparticles accumulate and reside in tumor sites, delivering chemotherapeutic agents.141 NanoGAG bioactive composites contain a number of different GAGs, thus having similar biochemical and mechanical properties with the natural PG, while their size is of the same order of magnitude as a PG; therefore, they serve as attractive vesicles in tissue regeneration as well as for targeted therapy and drug delivery systems.137 Several nano-GAG formulations with a wide range of biomedical applications have been described in the literature. Hep has been extensively studied as a drug delivery system, since it inhibits angiogenesis and metastasis, apart from its anticoagulant properties. The nanoencapsulated Hep analogues, mammalian unfractionated Hep and Styela plicataisolated Hep, exhibit significant antithrombotic effects with reduced hemorrhagic effects in vivo. The structure of ascidian Hep is more complicated, as it mainly consists of the disaccharide [α-L-IdoA(2)-1−4β-D-GlcN(S)(6S)-1]n, with differential sulfation patterns. These Hep nanoconjugates have significant antiproliferative and antimetastatic behavior in an aggressive breast cancer cell lines, therefore the potential therapeutic utilization of such analogues is crucial for pharmacological targeting.94,142,143 Covalent coupling of bioactive Hep to poly(ethylenimine)-coated multiwalled carbon nanotubes enhanced their blood compatibility and improved usage as building blocks for biomedical applications.144 Another important example includes Hep-conjugated silver nanoparticles by using AgNO3 and gold nanoparticles of HAuCl4 with chitosan to both the reducing and the stabilizing agent.145 These bioactive nanocomposites have significant antiangiogenic properties and anti-inflammatory properties without systemic hemostasis following injection.146,147 Interestingly, Hep nanoformations are used with gold nanoparticles and quantum dots in noninvasive biomedical imaging.148 Hepcoated nanoparticles composed of biodegradable hydrophobic cores [poly lactic-co-glycolic acid (PLGA)] and hydrophilic surface layers (Pluronic F-127), via a solvent-diffusion emulsion method and high-speed centrifugation, exhibited angiogenic bioactivity and supported the sustained release of M

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 2. Members of Acan/lectican/hyalectan family and link protein family (Crtl1/link protein1/HAPLN1).

G1 Domain. The N-terminal G1 domain, containing A, B, and B′ subdomains, binds to both HA and Crtl1 at the B−B′ stretch, and its binding affinity is enhanced by the A subdomain.174 This is in contrast to aggrecan, whose A subdomain of the G1 domain interacts with Crtl1.175 Each of B and B′ subdomain exhibits the link module, and the tandemly repeated orientation is required for adequate interaction with HA, whereas CD44 and TSG6 contain only one link module and exhibit adequate HA-binding activity.176 By HA-Vcan interaction, HA may require Vcan for incorporation into the ECM, or vice versa. Mouse embryonic fibroblasts (MEFs) with decreased expression of the mutant Vcan without the A subdomain exhibits decreased HA incorporation and altered HA-mediated signaling.177 In contrast, when HA expression is suppressed, Vcan incorporation is slightly affected. Recombinant human G1 domain exogenously added to cultured fibroblasts assembles HA on the cell surface into a cablelike structure.178 Vcan binds both HA and fibrillin, and when Vcan is cleaved into a G1-containing fragment, the fragments recruit HA into microfibrils.179 Therefore, it is likely that Vcan determines HA incorporation into the ECM and subsequent function. In cartilage, Acan is bound to HA; this complex is stabilized by Crtl1/HAPLN1, and these three molecules form the PG aggregate. Four members of the link protein family have been identified. They are renamed as HA and PG link proteins (HAPLNs) (Figure 2).180 This raises two questions: whether other Acan family members form the PG aggregate, and if so, which HAPLN is the partner of each PG. As expression of HAPLN2 and 4 is restricted to the nervous systems, like neurocan and brevican, they may interact with these brain-specific PGs, forming PG aggregates. Contrasting with these PGs, Vcan is transiently expressed during development and rapidly disappears; it may thus not form the stable PG aggregates. However, there are a couple of observations of Vcan aggregates. In the nodes of Ranvier, HAPLN2 is colocalized with Vcan V2 variant as well as brevican, and the nodes in HAPLN2-null mice lack a defined nodal pattern typified with a marked decrease of central nervous system (CNS) nerve conduction, suggesting the presence of HA, HAPLN2, and Vcan V2 and/or brevican in the nodes of Ranvier. As described in the section of heart development, Vcan is colocalized with Crtl1/HAPLN1, and either Vcan mutant or Crtl1/HAPLN1-null mouse exhibits

4. EXTRACELLULAR PROTEOGLYCANS 4.1. Versican: Structural Characteristics and in Vivo Functions

The first PG found was Acan, abundant in cartilage.164 Cartilage contains a large amount of Acan, which forms aggregates with both HA and cartilage link protein (Crtl1/link protein 1/HAPLN1).165 Using a method designed to isolate Acan, Vcan was isolated from bovine aorta166 and chick limb bud.167 To date, four members of the Acan/lectican/hyalectan family have been identified: Acan, Vcan, neurocan, and brevican (Figure 2). Dermacan, later identified in Zebrafish, is expressed only in fish.168 Even though Acan is present in cartilage and brain and both neurocan and brevican are restricted to nervous systems, Vcan is expressed rather ubiquitously and, therefore, is the prototypical member of this gene family.31,169,170 Cloning of Vcan revealed the presence of another EGF-like motif when compared with Acan.171 On the basis of this finding, it was named after its versatility. Studies have demonstrated at least two aspects of its versatility, splicing variants and characteristic expression patterns.169,172 This section is mainly focused on the structural characteristics and in vivo functions of Vcan. 4.1.1. Structure and Functional Domains. Vcan is a large CS/DS PG, mainly expressed in mesenchymal cells and located extracellularly. As a member of Acan/lectican/ hyalectan family, Vcan comprises four functional domains: HA-binding G1, TGF-β-regulating G3, CS chains, and an EGF-like motif (Table 1). The domains and their functions are presented below. Vcan splicing variants have recently been renumbered as Transcript Variant 1 (V0), TV2 (V3), TV3 (V1), and TV4 (V2). Both V0 and V1 are the major variants usually found in various mesenchymal tissues, whereas V2 variant is expressed in the adult brain.173 V3 protein has not been isolated from native tissues, as V3 without CS domains passes through anion-exchange chromatography. The core protein of Vcan contains both G1 and G3 domains at both termini, which is common among its variants and among aggrecan family members. Although the location and number of CS/DS chains attached to CSα and CSβ domains are still unclear, RNA splicing may generate Vcan with a different number of CS chains. N

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

ventricular septal defect (VSD), suggesting formation of Vcan aggregates in the atrio-ventricular canal (AVC).181 G3 Domain. The G3 domain contains two EGF-like motifs, a C-terminal lectinlike domain and complement regulatory protein (CRP)-motifs. Biochemical studies have identified a variety of ECM molecules that interact with the G3 domain, including tenascins, fibulins, fibrillins, and fibronectin, thus modulating cell motility, growth, and differentiation. Recent studies revealed the mechanism under which TGF-β is stored in the ECM. Latent TGF-β-binding proteins (LTBPs)-1, 3, and 4 bind TGF-β. LTBP-3 and 4 require fibrillin-1 to be incorporated into the ECM, whereas LTBP-1 does fibronectin. As the G3 domain interacts with fibrillin-1 and fibronectin, the G3 domain of Vcan may indirectly affect local storage and signaling of TGF-β. CS-Chains. Vcan has ∼20 putative GAG-attachment sites in CSα and CSβ domains. However, it is unclear how many chains are actually attached. CS of Vcan in mouse dermis contains approximately 35% DS units (IdoA-GalNAc4S). Previous studies have demonstrated that the V3 variant successfully restores the abnormalities, whereas chondroitinase ABC treatment has minimal effects on them, suggesting that the G1 and/or G3 domain is important rather than CS/DS chains. 4.1.2. In Vivo Roles in Development and Disease. Vcan can be isolated from aorta and brain, indicating that it serves as a structural macromolecule in the ECM. In contrast, it is transiently expressed, at high levels, during development, and the decreased expression or the absence of Vcan leads to impaired development of various mesenchymal organs and tissues, including heart, cartilage, and dermis. Analyses using various Vcan mutant mice have built up a concept that Vcan provides “provisional matrix”,31 which is followed by the construction of authentic ECM. While shifting from provisional to authentic ECM, Vcan disappears and is replaced by other molecules of constituting authentic ECM components. The provisional matrix generates the space for cell migration and differentiation and may increase local concentrations of growth factors and cytokines, both leading to development and maturation of mesenchymal organs. Heart Development. The first Vcan-mutant mice were generated by a gene trap technique. As they displayed heart defects, they were named hdf/hdf.182 Their Vcan gene has an insertion of a lacZ reporter construct in intron 7, which results in homozygous recessive embryonic lethality at approximately E10.5. The homozygote embryos are small and display severe cardiac defects along the anterior-posterior axis, including the absence of endocardial cushions and a highly dilated primitive atrium and ventricle. This is consistent with distinct expression patterns of Vcan in the developing heart.183 Generation of Vcan-null mice was attempted by inserting PGK-NeoR-polyA cassette into exon 3, which encodes the A subdomain of the G1 domain. The homozygotes, termed VcanΔ3/Δ3, express a decreased level of Vcan without the A subdomain by exonskipping and display various heart defects at different developmental stages (Figure 3).184 These mice with C57Bl/ 6 background die at E10.5 with heart defects like hdf/hdf, including dilated ventricles and lack of cardiac jelly. In contrast, VcanΔ3/Δ3 on a mixed genetic background survive longer, up to P0. Despite normal cardiac looping, these mice exhibit VSD. Their AVC cushion is much smaller, and the cells there exhibit endocardial and mesenchymal mixed characteristics with higher levels of proliferation. These observations suggest that

Figure 3. Wild-type (WT) and VcanΔ3/Δ3 embryos with the mixed genetic background of Balb/c. (A) Whole bodies of WT and (B) VcanΔ3/Δ3 embryos at E14.5 are shown. The VcanΔ3/Δ3 embryo is smaller when compared with the WT embryo. (C, E, and G) Sections of the WT and VcanΔ3/Δ3 (D, F, H) heart stained with HE at (C, D) E13.5 and (E−H) P0 are shown. LA, left atrium; LV, left ventricle; RA, right atrium; RV, right ventricle; and VSD, ventricular septal defect. Scale bars: (A) 1 mm, (C) 200 μm, and (G) 500 μm. Reprinted with permission from ref 184. Copyright 2012 Oxford University Press.

Vcan forms the PG aggregate with HA, which facilitates the formation of ventricular septa. Analysis of cardiac development revealed that both hdf heterozygotes and Crtl1/HAPLN1-null mice display VSD, indicating that the PG aggregate of Vcan with HA and link protein Crtl1/HAPLN1 is required to complete ventricular septa.181 Cartilage and Joint Development. Most bones, including long bones, are formed from cartilage tissue via endochondral ossification. Impairment of cartilage development leads to chondrodysplasia and dwarfism. Mesenchymal cells become condensed in limb bud regions, and after cell−cell interaction, they differentiate into chondrocytes, which express cartilagespecific molecules including collagen type II and Acan. Vcan is transiently expressed at high levels in mesenchymal condensation areas of cartilage primordium,167 and it rapidly disappears during chondrocyte differentiation.185,186 Cultured chondrogenic cells such as ATDC5 and N1511 attain similar expression patterns.186 In the initial stage of differentiation, Vcan is expressed at high levels. In turn, expression of Acan initiates and increases. By immunocytochemical analysis, Vcan O

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

deposited in the ECM is rapidly replaced by Acan. Micromass culture of hdf/hdf limb bud reveals the absence of collagen type II expression, indicating that Vcan is required for chondrocyte differentiation of limb bud cells.187 On the basis of these results, Vcan had long been believed to have important function, especially in the early stages of cartilage development. Analysis of Prx1-cre: Vcanflox/flox (Prx1-Vcan) mice clearly demonstrated the role of Vcan in cartilage development.188 These mice lack Vcan in mesenchymal condensation areas where Prx-1 is expressed. Although Prx1-Vcan mice are viable and fertile, they develop distorted digits. Histological analysis of newborn mice reveals hypertrophic chondrocytic nodules in cartilage, joint tilting, and a slight delay of chondrocyte differentiation. Interestingly, the lack of Vcan expression in limb bud impairs joint formation. The joint interzone is a future joint cavity, filled with mesenchymal cells. By immunostaining, whereas the control joint interzone shows an accumulation of TGF-β, concomitant with Vcan, that of Prx1-Vcan exhibits a decreased incorporation of TGF-β. In a micromass culture system of mesenchymal cells from limb bud, whereas TGF-β and Vcan are colocalized in the perinodular regions of developing cartilage in the control, TGF-β is widely distributed in Prx1-Vcan. Vcan facilitates chondrogenesis and joint morphogenesis by localizing TGF-β in the ECM and regulating its signaling (Figure 4). Dermal Development. It is reported that VcanΔ3/Δ3 MEFs in culture show premature senescence.177 Decreased levels of mutant Vcan together with the absence of the A subdomain, which dampens adequate HA-affinity, substantially suppress Vcan incorporation into the ECM, and alter HA-mediated signaling. The impact of Vcan in the ECM dynamism is also observed in vivo. When analyzed, VcanΔ3/Δ3 dermis at P0 exhibits an impaired ECM structure with reduced collagen type I and decreased cell density. When cultured dermal fibroblasts are analyzed, while the level of collagen deposition is similar, collagen biosynthesis significantly decreases in VcanΔ3/Δ3 fibroblasts as compared to wild-type (WT). VcanΔ3/Δ3 fibroblasts showed reduced TGF-β storage in the ECM and downregulation of TGF-β-Smad2/3 signal transduction, concomitant with downregulation of the early growth response (Egr) 2 and 4 transcription factors, which act downstream of TGF-β signaling.189 Mouse hair cycle consists of anagen, catagen, and telogen. Vcan is expressed at high levels in dermal papillae of both anagen and catagen, which suggests that Vcan plays roles in growth and maintenance of hair follicles.190 The VcanΔ3/Δ3 dermal papillae show decreased Vcan deposition as well as the number of hair follicles. RNAi targeted to Vcan efficiently suppresses the aggregative growth of dermal papilla cells in the hair follicle. HA is not colocalized in dermal papillae where Vcan is present. These observations support a role of Vcan in growth of hair follicles, which may exert independently of Vcan-HA interaction. Hair cycle-specific Vcan expression is also observed in humans,191 and aged-hair follicles exhibit remarkably decreased Vcan expression.192 Wagner Syndrome: VCAN Genetic Disorder. Wagner vitreoretinopathy (WGVRP; 143200) is an autosomal dominant vitreoretinopathy, originally described by Wagner and Johnston in 1983 as a middle lobe syndrome.193 Vcan gene mutations have been identified in this syndrome: an A-to-G transversion at the second base of the 3′-acceptor splice site of intron 7, a splice site mutation in intron 8, and three splice site mutations in intron 7, all of which are heterozygous. Analysis via qRT-PCR from patient blood samples reveals a significant

Figure 4. Conditional removal of Vcan in limb bud mesenchyme leads to limb deformities. (A) X-ray visualization of 8-month-old mice (a, b) and gross observations of 1-week-old mouse hind limbs (c, d). Prx1-Cre/Vcanflox/flox mice show distortion and shortening of digits (b, d), compared with Prx1-Cre/Vcan+/+ mice (a, c). (B) Histological analysis of newborn Prx1-Cre/Vcan+/+ (a, c) and Prx1-Cre/ Vcanflox/flox (b, d) hind limbs by hematoxylin and eosin staining. Prx1-Cre/Vcanflox/flox digits display joint tilting (arrows in b) and clefting (arrowheads in b), formation of hypertrophic chondrocyte nodules in proximal phalanges and delayed endochondral ossification (d). M, metatarsus; pp, proximal phalanges. Scale bars, 200 μm (a, b) and 40 μm (c, d). All histological sections (n ≥ 5) of Prx1-Cre/ Vcanflox/flox digits show similar abnormalities. Reprinted with permission from ref 188. Copyright 2010 American Society for Biochemistry and Molecular Biology.

and consistent increase in the V2 and V3 variants. Altered volumes and ratio of splice variants in vitreous body may cause abnormalities of Wagner vitreoretinopathy. It is intriguing that whereas mutations of Vcan in mice display a variety of abnormalities, human VCAN mutations are identified only in Wagner syndrome.194−196 Cancer Progression and Vcan. To date, there have been over three hundred reports demonstrating the close relationship of Vcan and cancer cell behavior: (i) Vcan expressed by cancer cells affects their own behavior, (ii) Vcan expression is a poor prognostic factor, (iii) Vcan expressed by cancer stromal cells regulates the structure of ECM and cancer microenvironment, and (iv) Vcan regulates cancer behavior via paracrine interactions. Several in vitro experiments using different cancer cell lines have shown that high Vcan expression could promote cell proliferation and invasion. The overexpression of the G3 domain can induce tumorigenesis and angiogenesis by elevating fibronectin and VEGF expression.197 TGF-β2 modulates glioma invasion by increasing Vcan V0/V1 expression and its cleaved fragment, termed versikine.198 V3 P

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

while the impact of Vcan expression on tumor cell behavior may vary depending on tumor types.

overproduction suppresses growth in vitro and tumorigenicity of melanoma cells, by altering the HA-CD44 interaction.199 Interestingly, the same group indicated Vcan as a possible proliferative, antiadhesive, and pro-migratory molecule.200 Independent studies using clinical samples have demonstrated that Vcan expression can be correlated with cancer progression, and its relevance is dependent on types and stages of cancer. Vcan expression in the cancer stroma is elevated in early stage prostate cancer, which predicts disease progression.201 Elevated Vcan expression in the tumor stroma could predict increased risk and rate of relapse in nodenegative breast cancer.202 Studies of endometrial carcinoma revealed that stromal Vcan expression is significantly higher in advanced-stage and high-grade cancers, lymph node metastasis, and ovarian metastasis, and epithelial Vcan expression is significantly higher in patients with lymph node metastasis and lymph-vascular space involvement.203 Low expression of the RhoGTP dissociation inhibitor 2 (RhoGDI2) is associated with metastatic disease in patients with muscle-invasive bladder cancer, and increased RhoGDI2 expression suppresses Vcan expression, which inhibits lung metastasis.204 Increased stromal Vcan is related to higher tumor recurrence rate and more advanced disease. Furthermore, Vcan expression in the gastric cancer stroma correlates with invasion, but not with other clinicopathologic features, such as age, gender, tumor size, tumor location, histologic grade, lymphatic metastasis, peritoneal metastasis, vascular invasion, nervous invasion, or tumor stages.205 The role of provisional matrix generated by Vcan and HA was demonstrated in a study of human leiomyosarcoma smooth muscle tumors. The tumor cells exhibit thick pericellular coats in culture, which could be reduced by inhibiting Vcan synthesis. Interestingly, inhibiting Vcan synthesis and reducing the pericellular matrices inhibit tumor cell growth in vitro and the ability of these cells to form tumors in vivo in a mouse model of leiomyosarcoma.206,207 Local depletion of Vcan expression by injecting cancer cells together with Cre-expressing adenovirus in Vcanflox/flox mice leads to faster tumor growth, indicating that stromal Vcan suppresses tumor cell growth. There, Vcan maintains the structure and volume of stroma by retaining collagen fibers and fibroblasts. Interestingly, Vcan and TGF-β are localized in a reciprocal manner, in contrast to the observations of their colocalization during development.208 Another study on cancer stroma reveals that Vcan promotes tumor angiogenesis. Tumors in hdf/+ mice show reduced growth with a lower capillary density and accumulation of capillaries at the tumor periphery. Vcan expressed in macrophages adjacent to capillary endothelial cells and cleaved Vcan fragments appear to facilitate tumor angiogenesis and tumor growth.209 Some mouse studies reveal that Vcan remotely controls cancer cells. Vcan expression in Lewis lung carcinoma cells activates the production of TNF-α and IL-6 through TLR2−TLR6−CD14 signaling in macrophages and myeloid cells, which facilitates metastatic tumor growth.210 In MMTV-PyMT transgenic mice with spontaneous breast cancer, the monocytic cells of a defined fraction in bone marrow that express Vcan migrate to the lungs, recruit metastatic breast cancer cells and promote mesenchymal to epithelial transition.211 As discussed in this section, there are cumulating studies demonstrating the role of Vcan in tumor progression. However, the way that Vcan could regulate tumor cell proliferation, invasion, angiogenesis, and metastasis may differ,

4.2. Decorin is a Multifunctional Proteoglycan

Decorin is a ubiquitously expressed CS/DS containing SLRP found in all type I and II collagen-rich tissues.212 Decorin was identified as a regulator of collagen fibrillogenesis, hence the eponym, required for preserving tissue integrity.213−216 In the following decades, it became clear that collagen fibrillogenesis was only the “tip of the proteoglycan iceberg” insofar as decorin is involved in a plethora of interactions and functions.217 Henceforth, decorin evolved as the archetype for understanding SLRP function. Decorin is a multifaceted signaling factor that underscores the importance of SLRPs in organismal homeostasis and pathobiology. This biological variety is seen in immunomodulation,218,219 calcium homeostasis,220 wound healing,221−223 keratinocyte function,224 diabetic nephropathies,225 fetal membrane signaling,226 obesity and type II diabetes,227 allergen-induced asthma and allergic inflammation,228,229 delayed hypersensitivity reactions,230 angiogenesis,231−234 hepatic fibrosis and carcinogenesis,235−237 myogenesis and muscular dystrophy,238,239 postmyocardial infarction remodeling,240 and mediating proper vertebrate convergent extension.241 Decorin is a biomarker for ischemic stroke,242 renal and pulmonary pathologies,243−245 as well as maintaining stem cell populations 246,247 and biomechanics.213,215,216,248,249 Recent work has focused on the role of decorin-impairing tumorigenesis and angiogenesis via highaffinity interactions with various cell surface signaling receptors expressed by the tumor and stroma, including EGFR, Met, and VEGFR2, respectively. The canonical roles of decorin as a panreceptor tyrosine kinase (pan-RTK) inhibitor exerting antitumorigenic and antiangiogenic effects on the tumor proper via suppression of potent oncogenes will be discussed in section 4.2.2. The decorin-evoked endothelial cell autophagy and tumor cell mitophagy, freshly identified neofunctions of decorin operative in the tumor microenvironment and tumor parenchyma will be also presented and discussed. This armamentarium ensures decorin as the guardian from the matrix.250 4.2.1. Structural Considerations of Decorin. Decorin is a constituent of an 18-gene member family and is decorated with a single N-terminal GAG chain of either DS or CS, 12 leucine-rich repeats (LRRs), and a C-terminal Ear domain (see section 2).251 The protein core of decorin is composed of tandemly repeated LRRs and are designated sequentially with roman numerals I−XII in a characteristic solenoidal architecture. LRRs are ∼24 AA in length with a conserved stretch of hydrophobic residues that form short β-sheets on the concave surface of the solenoid. These short β-sheets exhibit a parallel conformation with adjacent LRRs. On the convex surface of the solenoid, β-sheets are flanked by and intertwined with equally short β-strands joined by α-helices. At the N- and C-termini of each LRR are disulfide bonds that function as a cap. It is this structural variation of the caps that categorize the SLRPs into classes I−III versus classes IV and V.252 This fundamental LRR arrangement renders a functionally plastic binding interface capable of coordinating a multitude of protein−protein and protein−receptor interactions. This molecular feature is the reason for the widespread promiscuity decorin and related SLRPs.217 In functional analogy to the amino acid composition of zinc fingers found in transcription factors for sequence-specific DNA binding,253 individual AA Q

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 5. Schematic representations of the canonical and neoactivities of decorin. (A) Growth inhibition and angiostasis on cancer cells, (B) endothelial cell autophagy, and (C) tumor cell mitophagy. Please consult the text for additional information.

mechanistically implicates E-cadherin and oncogenic activin C.267,268 The magnitude of decorin-evoked RTK suppression is clearly evident in the Dcn null background of chemically induced models of hepatocellular carcinomas. Global genetic ablation of decorin results in many RTKs being constitutively activated.236 Loss of decorin permits anomalous, basal activation of several RTKs as determined by significant increases in phospho-Tyr signals.236 Indeed, transcriptomic profiling of hepatocellular carcinoma (HCC) shows decreased decorin expression, nicely complementing the genetic approaches elucidating the role of decorin in HCC.237 In the context of EGFR, Met, and potentially VEGFR2, monomeric decorin269 binds a narrow area that partially overlaps with the agonist binding cleft.270 Decorin binding promotes receptor dimerization, analogous to the natural ligands,271−273 followed by transient autophosphorylation of the unstructured intracellular cytoplasmic tails. Next is recruitment and activation of downstream effectors, internalization of the decorin/receptor complex, via caveosomes, and endosomal trafficking to lysosomes for degradation.261,274−276 Overall, this mode of action is considered dogmatic for decorin activity in the framework of tumorigenic and angiogenic RTK signaling. One notable exception to this paradigm must be addressed before detailing the downstream effects of decorin/EGFR and decorin/Met binding. This is the case of the IGF-IR/insulin system.264,277,278 Decorin directly binds IGF-IR, but unlike the three aforementioned receptors262 does not trigger internalization nor impair receptor complex stability.264,278 Alternately, decorin destabilizes IRS-1,264 thereby attenuating PI3K/Akt/ MAPK/Paxillin activation necessary for IGF-I induced mobility.279 Compounding the complexity of the decorin/ IGF-IR axis, decorin is a dichotomous IGF-IR ligand whose function is stringently tissue-dependent. In normal tissues, decorin is an IGF-IR agonist but functions as an obligate IGFIR antagonist in cancer.278,277 The concepts gleaned from the influence of decorin over the IGF-IR system can be applied to other SLRP family members in orchestrating crosstalk between

found within the LRRs confers specificity for the biological properties of decorin. This key property is exemplified by the ability of individual or tandem LRRs to bind specific decorininteracting molecules. As examples, LRR XII binds CCN2/ connective tissue growth factor (CTGF),254 whereas LRR V/ VI mediates the binding of decorin to VEGFR2,255 and the collagen-binding sequence, SYIRIADTNIT, residing in LRR VII, located on the concave region of the solenoid,256 mediates direct binding of decorin to collagen type I. Exclusively required for competent collagen fibrillogenesis, the covalently attached GAG chain plays a pivotal role252 and has been implicated in various connective tissue disorders, including Ehlers-Danlos syndrome257 and cancer.258 GAG chains that contain aberrant post-translational modifications or missing chains can severely impact the structural tenacity of the surrounding ECM.259 This is evident in the skin fragility phenotype in patients suffering from Ehlers-Danlos syndrome where nearly half of the secreted decorin is completely devoid of the N-terminal GAG chain.260 However, in the context of controlling intracellular signaling cascades via cell surface receptor binding, the GAG chain is entirely dispensable with the notable exception of a few cases.261 4.2.2. Tumor Suppression via pan-RTK Inhibition. An emergent property stemming from the individual structural determinants (e.g., LRRs), and overall solenoidal configuration of decorin, is the rather promiscuous nature of binding several cell surface receptors expressed by cells residing in the tumor microenvironment and the tumor proper. This property is critical for decorin to inhibit tumor progression and metastases.262 Decorin is an endogenous, soluble pan-RTK inhibitor, specifically sensitive toward “target rich” cells that overexpress EGFR, Met, and/or VEGFR2 (Figure 5).258,263,264 This specific trio of RTKs represent the most established and instrumental effectors for transducing signals necessary for decorin-mediated oncogenic and angiogenic suppression.250,262 There is a genetic interaction between loss of decorin and p53 as germline mutations in decorin and p53 accelerates lymphoma tumorigenesis.265 Moreover, Dcn−/− mice develop intestinal tumors when subjected to a high-fat diet266 and R

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Downstream of Met antagonism, two potent oncogenes, βcatenin and Myc, are targeted for degradation via the 26S proteasome (Figure 5A, right).290 Suppression of the MYC locus in conjunction with phosphorylation-dependent protein degradation of Myc at Thr58, well-within the N-terminal degron domain, derepresses the CDKN1A locus via loss of the AP4 repressor.290 A second mechanism of CDKN1A induction is discussed in section 4.2.3. Moreover, decorin silences βcatenin signaling in a noncanonical manner that is independent of Axin-1/DSH/GSK-3β.290 In this mode of action, β-catenin is phosphorylated, not for increased protein stability, but for proteasomal degradation291 (Figure 5A, right) in a manner consistent with direct phosphorylation of β-catenin by an RTK, such as Met.292−295 Noncanonical governance of β-catenin by decorin may underlie the mechanism of intestinal tumor formation following decorin ablation. Intestinal epithelium turnover and maturation is propelled by β-catenin.296 Met is constitutively activated in colon carcinoma and directly influences β-catenin signaling.297 Given that global loss of decorin derepresses multiple RTKs in two distinct chemically induced models of HCC,236 malignant transformation of the intestinal epithelium (or other solid malignancies) is triggered by interfering with the Met/β-catenin axis. Met targeting by decorin is the primary node for angiogenic suppression in cervical and breast carcinomas.250 Similar to decorin/EGFR, Met signaling decreases HIF-1α independent of ambient oxygen tensions (Figure 5A, right).250 Correspondingly, VEGFA synthesis is compromised in several in vitro studies utilizing primary endothelial cells, MDA-MB-231 triplenegative breast carcinoma cells, and in vivo as demonstrated with HeLa tumor xenografts.250 MMP-2/9 (gelatinase A and B, respectively),298−307 which liberate VEGFA tethered to matrix components, are suppressed by decorin,250 further contributing to angiostasis. Intriguingly, MMP-9 processes the angiostatic Multimerin-2,308 which binds CD39 on the surface of activated endothelium.309 Decorin evokes the simultaneous expression and secretion of TIMP3 and TSP-1, potent antiangiogenic factors (Figure 5A, right).250 Decorin triggers the rapid secretion of TSP-1 from MDA-MB-231 cells in an EGFR-dependent manner by attenuating the RhoA/ROCK1 signaling.233 Given the powerful antiangiogenic activity of TSP-1 and its involvement in pathophysiology,310−315 it is likely that decorin subsumes a protective role against cancer growth and metabolism. As decorin suppresses HIF-1α, it is possible that decorin derepresses TSP-2 expression.316 A second mechanism for TSP-1 was discovered; consult section 4.2.4. 4.2.3. Decorin-Evoked Catabolism: Core of the Proteoglycan Iceberg. A second breakthrough discovery in elucidating the role of decorin in vivo derived from a highresolution preclinical genomics screen that evaluated novel decorin-regulated genes.317 Triple-negative breast carcinoma orthotopic xenografts were established and treated systemically with decorin, and expression profiles were obtained.317 In contrast to conventional microarrays, this platform was designed for codetection of species-specific genes modulated within the host stroma (Mus musculus) and those being regulated in the tumor xenograft itself (Homo sapiens).317 Authentication of the significantly (and differentially) modulated gene targets revealed a small subset of decorin-regulated genes, exclusively within the murine tumor microenvironment.317 Intriguingly, minimal transcriptomic changes occurred in the human tumor cells.317 The transcriptomic

EGFR/IGF-IR in estrogen receptor-positive breast cancer.280,281 Embracing the concept of binding promiscuity, cell surface receptors with specific structural motifs, specifically members of the IgG superfamily, may provide insight and further discoveries of additional RTKs targeted by decorin.217,270,272 Indeed, the ectodomains of EGFR, Met, and VEGFR2 all contain multiple IgG motifs.282,283 Mechanistically, decorin binding may promote a combinatorically different phosphorylation signatures than the pattern generated by agonist binding such as those elicited by TGF-α, EGF, HGF/SF, VEGFA, VEGFB, etc. EGFR: Original Story. The original dogma-shattering concept of decorin-evoked RTK-antagonism was birthed following the discovery that EGFR is a decorin target284 and that decorin functions as an endogenous ligand to modulate receptor activity.265 Evaluating decorin in A431 orthotopic tumor xenografts revealed significant suppression of tumorigenic growth by targeting EGFR in vivo.285 Mechanistically, decorin indirectly inhibits HER2/ErbB2 heterodimerization,286 via titration of signal-competent ErbB1/ErbB2 heterodimers.262 In the central nervous system, decorin binds and represses ErbB4/STAT3 signaling,287 suggesting tissuespecific effects. Decorin triggers transient activation of ERK1/2 signaling paradoxically dependent on the EGFR tyrosine kinase288 concurrent with a regulated burst of cytosolic Ca2+.220 As a recurring theme following decorin/receptor ligation is a wave of positive receptor signaling despite a >50% reduction of total EGFR to evoke p21WAF1 induction with concomitant proteolysis of pro-caspase-3 into active caspase 3 (Figure 5A, left).288 Activation of this pathway promotes cell cycle arrest and induces apoptosis, respectively. Imperative for the protracted function of decorin, decorin/EGFR complexes are shuttled into caveolin-1 coated pits.234 Phosphorylation of specific residues are required for caveolin-1/EGFR association289 and subsequent internalization for endocytic degradation. Shuttling of EGFR into caveosomes shortens the duration and frequency of EGFR signaling in contrast to active ligand which sorts EGFR into clathrin-coated pits. This leads to endosomal recycling and, ultimately, to repopulation of the cell surface with activated EGFR for additional rounds of signaling, conducive for sustained tumorigenic and angiogenic processes. Decorin Meets Met-Digging Deeper into the Proteoglycan Iceberg. A recurring theme of decorin oncosuppression involves transient activation of the decorin/receptor complex.250 Utilizing a phosphotyrosine RTK array as an unbiased discovery platform, a second RTK, Met, sobriquet as the HGF/SF receptor, is activated by decorin.272 Met is the primary decorin receptor expressed by the tumor and is ultimately responsible for relaying signals for antitumorigenic, antiangiogenic, and pro-mitophagic programs (see section 4.2.5).262,272 Decorin has a tighter binding affinity for Met than EGFR, (Kd ∼ 2 vs 87 nM, respectively).272 Mechanistically similar to EGFR, decorin/Met complexes are transited from the cell surface into caveolin-1 positive endosomes post recruitment of c-Cbl to Met at Tyr1003 phosphorylation (Figure 5A, right), a post-translational modification favored by decorin.272 The trimeric complex of decorin/Met/caveolin-1 ensures presumed delivery to lysosomes and a cessation of oncogenic signaling. This mechanism is in congruence with the fate of decorin/EGFR complexes in contrast with HGF/Met, EGF/EGFR, or TGF-β/EGFR.290 S

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

a partial agonist as signaling through the RTK kinase domain is required to transduce the information encoded by the LRRs of decorin. Downstream of stimulated VEGFR2, decorin differentially regulates signaling by activating pro-autophagic multiprotein complexes including AMPKα and Vps34 positive nodes, while concurrently attenuating, in a protracted fashion, the antiautophagic effectors, PI3K/Akt/mTORC1 (Figure 5B).344 Overall, decorin triggers excessive and protracted autophagy as measured by the degree of Beclin 1/LC3 dually positive autophagosomes (Figure 6), by increasing the rate and number

signature ushered in the advent of a new functional paradigm in understanding decorin activity. Decorin reprograms the tumor microenvironment and tumor parenchyma toward a pro-catabolic state by evoking endothelial cell autophagy and tumor cell mitophagy. This discovery came full circle as autophagic stimuli regulates Dcn itself. 4.2.4. Decorin Evokes Endothelial Cell Autophagy in a Peg3-Dependent Manner. The high-resolution genomics screen identified several decorin-inducible genes within the tumor stroma.317 More focus is given on the maternally imprinted C2H2 Krupple zinc-finger-containing transcription factor, PEG3.317−321 Peg3 regulates stem cell progenitor development,322,323 mediates p53-dependent apoptosis of myogenic and neural pedigrees,324−328 and determines maternal/paternal behavioral.329,330 Considered a tumor suppressor gene,331 PEG3 is frequently lost via promoter hypermethylation and/or loss of heterozygosity332−335 in cervical and ovarian cancer. Peg3 is unique among the decorin inducible-tumor suppressor genes, as Peg3 was exclusive to the stroma and shares the same noncanonical, GSK-3β independent suppression of the Wnt/β-catenin pathway as discovered with decorin.336 Peg3 was evaluated within endothelial cells, a surrogate for the tumor microenvironment, the major cell type coordinating tumor neovascularization. These cells are sensitive to nanomolar concentrations of soluble decorin, which suppresses VEGFA, a major endothelial cell survival factor.234 Serendipitously, Peg3 relocalizes to intracellular structures reminiscent of autophagosomes.273 Subsequent studies confirmed that subcellular localization patterns of Peg3, in the presence of decorin, overlap with those of established autophagic markers such as Beclin 1 and LC3,337,338 thereby illuminating a heretofore unidentified autophagic protein. Differential binding of Peg3 to LC3-II and an augmented binding of Peg3 to Beclin 1 following autophagic stimulation by decorin were found biochemically (Figure 5B).273 Peg3 is required for decorin-evoked BECN1 and MAP1LC3A expression and is responsible for maintaining Beclin 1 under basal conditions (Figure 5B).273,339 BECN1 was confirmed as a direct target of Peg3,340 as de novo expression of Peg3 was sufficient to drive BECN1 expression and protein as assessed by promoter luciferase assays and immunoblotting in a dose-dependent manner, respectively.340 Full length Peg3 was required for this process as loss of the N-terminal SCAN domain or only the zinc fingers was not sufficient.340 Peg3 is sufficient to drive increased LC3 puncta formation and augments autophagic flux,340 by assaying LC3-II levels. The same requirement of full-length Peg3 was also required to drive TFEB expression in a dose-dependent fashion (see below).341 Decorin drives a pro-autophagic signaling program under the influence of Peg3, Beclin 1, and LC3 while combinatorically precluding Bcl-2 from this complex, a known autophagic inhibitor.342 Unlike other autophagic stimuli such as nutrient and/or amino acid deprivation, mTORC1 inhibition by rapamycin or Torin-1,338 decorin induces autophagy, as per the modus operandi of decorin, downstream of a cell surface RTK, in full nutrient conditions. Decorin binds VEGFR2, the King Kong of receptors for endothelial cells, for autophagic induction (Figure 5B).273 Pharmacological inhibition with SU5416343 abrogates the RTK-governed autophagic response, suggesting that decorin requires the VEGFR2 tyrosine kinase for autophagy.273,339 It appears the original conceptual advance of decorin functioning as an unbridled pan-RTK is evolving as

Figure 6. Decorin and endorepellin evokes Beclin 1/LC3 dually positive autophagomes in endothelial cells. (A) Differential interference contrast microscopy of autophagosomes immune-stained for Beclin 1 (green) and LC3 (red) following decorin stimulation. (B) Representative fluorescence micrographs of porcine aborting endothelial cells (PAE) cells stably transfected with GFP-LC3 and immune-stained for Beclin 1 (red) following treatment with endorepellin. Nuclei were visualized with DAPI (blue). Scale bar ∼ 10 μm.

of autophagosomes, for bulk degradation.345 Therefore, decorin must regulate the terminal end point of autophagy, that is autophagosomal fusion with lysosomes, to sustain this increased autophagic flux, both in content and over time. Thus, focused attention was given to TFEB, a master transcription factor for lysosomal biogenesis and an emerging factor in the nuclear control of autophagy.346 TFEB is an autophagic transcription factor347−349 that specifically binds CLEAR-box sequences present in the regulatory regions of lysosomal genes, for increased expression.350 TFEB is normally inactive under nutrient-rich conditions and is directly phosphorylated by mTORC1 and sequestered via 14−3−3 scaffolding proteins. Following an autophagic signal such as decorin, rapamycin, Torin-1, TFEB is rapidly dephosphorylated by calcineurin for nuclear accumulation and transcriptional activation of lysosomal genes.350 As decorin attenuates mTORC1 over a long period of time, TFEB is presumably dephosphorylated and its expression induced (Figure 5B).341,344 Mechanistically, decorin causes Peg3 nuclear translocation or following other autophagic stimuli340 and is situated upstream of TFEB (Figure 5B). Peg3 is necessary and sufficient for driving TFEB induction. Peg3 depletion is sufficient to inhibit decorin-mediated TFEB expression at both the mRNA and protein level. Moreover, increasing concentrations of Peg3 drive TFEB expression to comparable levels achieved by decorin. Inhibiting key effector kinases, such as VEGFR2 or AMPKα with SU5416 or Compound C (Dorsomophin), respectively, results in a complete blockade of TFEB, thereby placing TFEB as the most downstream molecule, to date, within the novel decorin/VEGFR2/ T

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

AMPKα/Peg3 pro-autophagic axis (Figure 5B).350 This is the first work that connects a member of the SLRP family to lysosomal biogenesis, a critical process for sustained autophagy and perhaps other lysosomal processes involving TFEB.351 The enhanced formation of dually positive Beclin 1/LC3 autophagosomes by decorin is abjectly abolished when TFEB expression is silenced, delineating a pivotal role of TFEB in the induction of crucial autophagic genes and in the structural formation of autophagosomes in endothelial cells. Nutrient status is a critical regulator of autophagic activity, and while aberrations in both autophagy and metabolism are implicated in tumorigenesis, these processes have also been implicated in many cardiac disorders. As a matter of happenstance, decorin mRNA and protein levels are induced upon fasting in cardiac tissue downstream of mTORC1 inhibition in vivo.352,353 This finding posits decorin as the first SLRP whose expression is induced by starvation signals. Biglycan, the structurally most related SLRP to decorin, was unchanged by the same autophagic cue, suggesting a specific response for decorin in cardiac tissue. Functionally, lack of decorin resulted in aberrant cardiac autophagy when compared to wild-type animals. Several autophagic markers, including MAP1LC3A did not respond to starvation in Dcn−/− mice, whereas a significant induction of MAP1LC3A is found in fasted wild-type animals of the same genetic background. Using green fluorescence protein (GFP)-LC3 reporter mice, a significant increase in decorin immunostaining as well as in GFP-LC3 puncta was observed compared animals fed ad libitium. However, compound animals do not show any increase in GFP-LC3 puncta as for starved animals compared with control fed littermates. A number of important genes are upregulated in the heart of wild-type animals under poor-nutrient conditions. Intriguingly, Cdkn1a, the gene encoding p21WAF1, is significantly increased in wild-type starved hearts but not in decorin knockout (KO) mice, indicating that decorin is necessary in vivo for p21WAF1 induction upon fasting. It is unknown whether this is the result of decorin/EGFR interactions or represents a novel pathway. Concomitant with autophagic activation, decorin significantly impairs capillary morphogenesis.232,250,273 Decorin triggers the rapid release of TSP-1 from breast carcinoma cells in a ROCK1-dependent manner. A second pathway for TSP-1 synthesis and secretion has emerged.340 Endothelial cells stably expressing Peg3 exhibited impaired wound healing and evasion from MatriGel due to increased production and secretion of TSP-1 into the media.340 Neutralizing antibodies against TSP1 were sufficient to override this profound wound healing inhibition driven by Peg3.340 Emerging evidence suggests that RTKs such as EGFR inhibit autophagy.354,355 Decorin is a partial RTK agonist that ultimately degrades cell surface RTKs. Therefore, autophagic regulation, as a result of RTK binding and inhibition, represents an innate property of decorin. 4.2.5. Decorin Evokes Tumor Cell Mitophagy in a Mitostatin-Dependent Manner. Decorin directly influences catabolic programs within the tumor proper itself as well. Respiratory complex turnover was prolonged, up to 24 h, and mitochondrial DNA (mtDNA) depletion, key signs of mitochondrial autophagy (mitophagy) (Figure 5C).356 Activation of this fundamental pathway may reconcile the canonical effects of decorin (section 4.2.2) with the paradigmatic discovery of matrix-mediated catabolism for impeding angiogenesis and tumorigenesis. In a mechanism analogous to VEGFR2, decorin requires Met for proper mitophagy and

respiratory chain turnover in triple negative and luminal breast carcinoma cells.356 Conceptually, autophagic/mitophagic activation requires cell surface RTKs and the inherent kinase activity of the cognate receptor. This form of autophagic and mitophagic induction is defined as noncanonical insofar as this requires cell surface RTK and occurs in nutrient-rich conditions. Similar to endothelial cell autophagy, tumor cell mitophagy is dependent on a central decorin-inducible tumor suppressor gene (Figure 5, panels B and C). Originally known as Ts12q (tumor suppressor at 12q), the protein encoded by TCHP (chromosome 12q24.1) has been eponymously renamed as mitostatin, for mitochondrial protein with oncostatic activity.357 Mitostatin was identified as a decorin-inducible gene via subtractive hybridization of cDNA libraries with probes obtained from vehicle-transferred or decorin-transfected cells.357 Several decorin-inducible genes were discovered from the cDNA libraries. Mitostatin is broadly expressed in normal tissues and is evolutionarily conserved across multiple species. Similar to Peg3, mitostatin is frequently lost and/or mutated at highly conserved residues throughout the open reading frame in bladder and breast carcinomas,357,358 supporting its role as a tumor suppressor. Mitostatin staining reveals a highly punctate pattern, primarily localized to the mitochondria358 and specifically enriched in mitochondrial-associated membranes (MAMs).359 MAMs are highly specialized connections formed by the juxtaposition of endoplasmic reticulum and mitochondria, where mitostatin directly binds mitofusin 2.359 Overexpression of mitostatin results in overt ultrastructural changes in mitochondrial architecture including loss of mitochondrial matrix, abnormal cristae, and a swollen appearance.357 Upon decorin/Met interactions, PGC-1α360 is mobilized and stabilizes TCHP mRNA for mitostatin accumulation.356 PGC1α binds TCHP mRNA via the C-terminal RNA recognition motif (RRM) (Figure 5C).356 Ablating the RRM domain or preventing proper methylation of arginine residues within the RRM via PRMT1 depletion abrogates TCHP mRNA stabilization and mitostatin accumulation.356 Elucidating this pathway revealed a unique and unlikely molecular cooperation between a mitophagic effector and a proto-oncogene. PGC-1α mediates oxidative metabolism361 in a subset of aggressive melanoma classified by increased mitochondrial capacity.362 Silencing mitostatin precludes turnover of various respiratory chain components, decreased TFAM and mtDNA content, VDAC1 clearance, and fragmentation of the mitochondrial network356 all established mitophagic markers (Figure 5C).363 Mitochondrial network fragmentation via decorin (Figure 5C) is consistent with the results of mitostatin overexpression. Assessing Su9-GFP, a well-established mitochondrial matrix protein, laser confocal microscopy and conventional immunoblotting techniques reveal a decrease in total Su9-GFP (Figure 5C), suggesting clearance via the autophagolysosomal system. Prior to fragmentation and aggregation of the tubular mitochondrial network, decorin triggers the mitochondrial membrane potential (ΔΨm),356 with a magnitude comparable to the established protonophore, FCCP (Figure 5C). Loss of ΔΨm across the mitochondrial membranes is an early harbinger of mitochondrial dysfunction, ATP depletion, and an effective signal for Parkin recruitment and subsequent turnover.364,365 Decorin-evoked ΔΨm may manifest downstream of the decorin/EGFR interaction from increased U

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

cytosolic Ca2+ levels.220 As mitostatin is localized to MAMs and interacts with mitofusin 2, it may permit an efflux of stored Ca2+ from the ER directly to the mitochondrial matrix as the starting impetus for decorin-evoked mitophagy (Figure 5C, pathway question mark). Decorin may activate the Parkin-mediated arm of mitophagy and respiratory chain turnover366 in a mitostatin-dependent manner (Figure 5C, pathway question mark). Structurally, mitostatin is predicted to contain up to five coiled-coil domains that may act to recruit and/or function as a coreceptor for Parkin/mitophagy receptor interactions at the OMM. Intriguingly, Beclin 1, a tumor suppressor and essential autophagy mediator354,367−369 is also composed of coiled-coil domains essential for Vps34 binding370 and Vps34 complex incorporation.371 Mitostatin may have functionally similar roles as a mediator of protein complexes relevant for mitophagy. Alternatively, mitostatin may indirectly stimulate inherent PINK1 kinase activity for proper targeting, ubiquitin activation,372,373 and/or Parkin-mediated ubiquitination of mitochondrial proteins.374−376 Collectively, Met signaling is necessary for mitophagic induction, in a mitostatin-dependent manner, within the tumor parenchyma of breast and prostate carcinomas.377 Mitophagy, in conjunction with endothelial cell macroautophagy, may represent a nexus for integrating decorin/ RTK outputs. These research lines coalesce around the conceptual advance that decorin-evoked endothelial cell autophagy and tumor cell mitophagy via RTK suppression serve as the molecular mechanism for suppressing tumor neovascularization and inhibiting tumor growth.

lumican expression has been associated with a poor outcome in breast cancer and may limit tumor invasion.388 Lumican inhibits primary melanoma tumor growth389 and lung metastasis in vivo.390 Melanoma cells adhere to lumican,391 contributing to cytoskeleton remodeling and inhibition of migration.392,393 The above data were confirmed in an in vivo mice model, where lumican inhibited the tumor growth and as a protagonist of ECM coherent assembly.394 Integrin α2β1 is a direct lumican receptor on melanoma cells.395 Lumican expression is preferentially located at the margin of the tumor in stromal dermal fibroblasts.396 Moreover, lumican was demonstrated to inhibit endothelial cell invasion, angiogenic sprouting, and vessel formation in mice.390,397,398 Lumican was detected in the cytoplasm of cancer cells and/ or stromal tissues adjacent to cancer cells in lung cancer.399 Lumican expression in cancer cells is correlated with pleural invasion and larger tumor size in lung adenocarcinoma. It was associated with the formation of a keratinized pattern in squamous cell carcinoma. Stromal lumican expression correlated with vascular invasion but its expression level in cancer cells did not correlate with patient prognosis. Lumican was strongly expressed in cancer cells, in acinar and islet cells in chronic pancreatitis-like lesions adjacent to the pancreatic tumor.385 Moreover, it was localized in fibroblasts and on collagen fibers close to cancer cells. Lumican stimulates growth but inhibits invasion of human pancreatic cancer cells.400 In colorectal cancer, lumican is synthesized by cancer cells, fibroblasts, and epithelial cells adjacent to cancer cells suggesting that it regulates the growth of human colorectal cancer.401 The migration of human colon cancer cells was shown to be regulated by lumican overexpression.402,403 Lumican also modulates progression of osteosarcoma.404 Altered MMP-14 activity proved to be one mechanism of action for the antitumorigenic properties of lumican.384,405,406 Indeed, it alters cell migration of Snail-transfected B16F1 cells.407 Synergism between MMP-14 and integrins is a hallmark in tumor invasion and angiogenesis.397,407 Lumican significantly inhibits cell proliferation, migration, and invasion in breast cancer models of different estrogen receptor status by EMT reprogramming and changes in MMP expression.93 Altogether, these data suggest that lumican may regulate tumor progression by either direct interactions with ECM molecules or modulation of membrane receptors379,384 and MMP14.406−408 Lumican-Derived Peptides Mimic the Antitumor Effect of Lumican. The human lumican transcript (1014 bp) encodes a protein of 338 AA. It includes an 18 AA signal peptide allowing its secretion into the ECM. Three major domains were identified: a negatively charged N-terminal domain containing sulfated Tyr and Cys residues, a central core containing 9 LRRs, and a C-terminal domain of 66 AA harboring two conserved Cys residues and two LRRs.381,409 Human lumican exhibits 11 LRRs motifs (LxxLxLxxNxL), characterized by a common molecular architecture adapted to protein−protein interactions.409 The central part of the lumican sequence is highly conserved as demonstrated by sequence alignment of different species. Four potential sites for the substitution by Nlinked KS or oligosaccharides are situated at position 87, 126, 159, and 251 of the core protein of human lumican.409−412 It exhibits secondary and tertiary structures, which are characteristics of the SLRPs.413,414 Lumican interacts with the transforming growth factor-β receptor 1 (TGF-βR1 or ALK5), independently of its glycan moiety.415 The last 13

4.3. Lumican Structure−Function Relationship: Key Findings by Molecular Modeling

Lumican (a member of SLRP family) was first identified as a major PG of the cornea, with a 38 kDa core protein.378−380 Lumican is expressed in the ECM of different tissues. In some tissues, such as skin, lumican may exist in glycoprotein form, being substituted with short oligosaccharides or poorly sulfated/unsulfated polylactosamine chains rather than KS.381,382 Its structures vary not only according to the tissue but also during aging.383 This section focuses on the various molecular forms of lumican and the corresponding biological effects on tumor progression, with a particular focus on the lumican-derived peptides interacting with MMP-14. 4.3.1. Lumican Regulates Tumor Progression in a Tissue-Specific Manner. There are four structural forms of lumican: the core protein of 38 kDa, which is nonglycosylated, a 57 kDa form possessing short N-linked oligosaccharides, a form in which the oligosaccharides are substituted with polylactosamine chains, and a form in which sulfation of the polylactosamine chains occurs to give KS.384 Here, the structural variability and the respective biological effects will be addressed. Lumican Structural Heterogeneity Leads to Heterogeneity in Its Control of Tumor Progression. Lumican role in tumor progression was reported to be either positive or negative. These discrepancies may be related to the variability of its structure (glycoprotein or KSPG) and the different affinities of its forms for integrins and MMPs. Lumican inhibits melanoma progression in vitro and in vivo.384 However, its expression in pancreatic cancer correlates with an advanced stage of invasion.385 In mammary cancer, lumican is not correlated with prognostic factors.386,387 In contrast, reduced V

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

AA of the lumican core protein (named LumC13 or lumikine) mimics the effects of the complete lumican molecule. Recently, LumC13 was demonstrated to form a stable complex with ALK5. This peptide significantly promoted corneal epithelium wound healing.416 A sequence of 17 AA (named lumcorin) within LRR9 was shown to reproduce the inhibitory effect of lumican on cell migration in vitro.417 The aligment of lumican, fibromodulin, decorin, and biglycan revealed that the AA corresponding sequence of lumcorin is highly conserved. However, only lumcorin was able to significantly inhibit recombinant MMP-14 activity.405 In addition, our group identified a 10 AA minimal active sequence (L9M) within the lumcorin sequence.417 Lumcorin and L9M inhibited melanoma cell growth in a similar manner.405 However, L9M peptide was more efficient in decreasing melanoma cell migration as compared with lumcorin. This effect might be explained, at least in part, by the slight difference in inhibiting the MMP-14 activity by lumcorin and L9M.405 The central fragment of decorin, LRR5, was also more active in the inhibition of angiogenesis than the full-length LRR.418 Similarly, the central 12 AA region of decorin LRR5 more efficiently inhibits angiogenesis than full length LRR5. Several fragments from ECM components were demonstrated to regulate tumor progression. The NC1 domains of the α3 (III) chain and α4 (IV) chain of collagen type IV were described to decrease tumor growth.419,420 Other peptides derived from connective tissue glycoproteins, like anastelin, may also decrease tumor growth.421 Thus, lumican and its derived peptides, are able to inhibit melanoma cell migration by decreasing MMP-14 activity.405,406 However, in silico modeling approaches will be necessary to clarify the interaction of lumcorin and L9M peptide with MMP-14. 4.3.2. Lumican Modeling Reveals Specific Interactions with Its Mediators. Lumican decreases melanoma cell migration by inhibiting MMP-14 activity by direct interaction with its catalytic domain.406 Using modeling approaches,408 the direct interaction of lumican to the catalytic domain of MMP14 activity is illustrated in Figure 7. Similarly, the modeling of lumican-derived peptides is starting to reveal their refined interactions with their different mediators. In this section, recent data associated with (i) the modeling of lumican and lumican-derived peptides and (ii) the understanding of the interaction of MMP-14 with lumican-derived peptides will be discussed. In order to study interactions between lumican, its derived peptides, and other proteins of the ECM, it is crucial to rely on 3D structure of the molecular partners. Indeed, the modeling of proteins requires the knowledge of a set of coordinates describing the positions of the atoms within the protein. Two main experimental techniques allow researchers to access these types of data: X-ray crystallography and NMR. Even though the data accessible through the Protein Databank (https:// www.rcsb.org/) are growing incredibly fast, no experimental structural information is available for human lumican core protein. Alternatively, it is possible to access 3D structures using methodologies such as homology modeling reconstruction. This technique was used in theoretical studies where the effects and interactions of lumican were screened. Thus, Zeltz and colleagues417 proposed a 3D structure of lumican deduced from bovine decorin (PDB 1XKU) and biglycan (PDB 2FT3) using the Swiss-Pdb Viewer software.422 On their model of the core protein, they were able to identify the position of the

Figure 7. Schematic representation of human MMP-14 transmembrane protein dimer interacting with human N-glycosylated lumican. Surface representation is used to depict the proteins and lumican PTMs. The cellular membrane is shown with sand-colored spheres, the MMP-14 dimer, lumican, and N-glycosylations are colored in red, blue, and orange, respectively. The region of lumican corresponding to lumcorin is highlighted in cyan. MMP-14 dimer was built using the crystal structures of the catalytic domain (PDB no. 1bqq) and the hemopexin-like domain (PDB no. 3c7x). For illustrative purpose, the transmembrane region was modeled using integrin αIibβ3 structure (PDB no. 2k9j). The lumican molecule was modeled after fibromodulin crystal structure (PDB no. 5mx0), and biantennary glycans were grafted on its surface. Lumican was shown to be involved in the inhibition of MMP-14 activity via a direct binding to its catalytic domain (such as the one depicted in this schematic representation): this direct interaction regulates cell motility, invasion, and metastasis.

LLR9 domain, which contains lumcorin. A few years later, Yamanaka and colleagues415 built a 3D model of the lumican core protein using MODELER 9v1 and the structure of bovine decorin (PDB 1XCD) as a template. This model of lumican was then used in docking experiments (performed with the HEX 4.5 software) in order to decipher the possible interactions with various receptors and predicted TGF-βR1 as the most probable partner. This software assumes that the molecules (ligand and target) are rigid, and since this consideration would be rather extreme and wrong for peptides, their intrinsic flexibility was evaluated indirectly using classical molecular dynamics simulations. It was then possible to extract the most relevant conformations with dedicated clustering protocols. One can consider that the 3D models of lumican obtained in these studies are trustworthy since the degrees of identity and homology are higher than the 30% threshold considered as a good guaranty for reconstruction through homology modeling. In addition, both models of the lumican core protein display the arch structure, which is the signature of SLRPs as predicted by Kajava using molecular modeling analysis.423 Recently, the authors who proposed the second model of lumican used an in silico approach,416 in order to unravel the crucial lumican AA sequence in the formation of the stable lumican/TGF-βR1 molecular complex. A set of eight peptides with a length of 13 residues and designed according to the C-terminal region of lumican were investigated following a comparative protocol that combined molecular docking experiments, molecular dynamics simulations, and free-energy decomposition analysis. One motif was highlighted as essential in the molecular complex formation and confirmed experimentally with wound healing assays: the EVTL motif. W

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 8. Schematic representation of endorepellin activity on endothelial cells. (A) Growth inhibition and angiostasis, (B) autophagy, and (C) autophagic gene signature. Please consult the text for additional information.

reveal a behavior like an early response gene whose transcription is blocked by interferon-γ (IFN-γ)428 and induced by TGF-β.429 Perlecan is localized to vascular basement membranes and pericellular spaces.430 Similar to other HSPGs,431 perlecan has a modular architecture which facilitates a multitude of regulatory processes including cell adhesion,432 articular cartilage formation,433 endochondral ossification,434,435 osteoarthritis and mast cell mediatedinflammation,436,437 endocytosis,438 peripheral node assembly,439 lipid catabolism,440 and thrombosis and cardiovascular development.441−443 Despite this myriad of homotypic and heterotypic interactions, perlecan is a fundamental rheostat of developmental and cancer angiogenesis258,443−452 and more recently of autophagic regulation.453,454 Perlecan maintains the delicate balance of factors involved in angiogenesis via intramolecular coordination with the Cterminal domain known as endorepellin,37,455−457 which interacts with the pleiotropic growth factor, progranulin.457,458 Endorepellin possesses intrinsic angiostatic properties447 and is antithetic in function to the triad of dynamic459 N-terminal HS chain coreceptors and growth factors repositories,432,446 which also have roles in branching morphogenesis.460 Endorepellin is cleaved from perlecan and detected in vivo436 and can be further processed by BMP1/Tolloid-like metalloprotease461 and cathepsin L.462 Structurally, endorepellin is composed of three laminin-G-like domains (LG1−3) each separated by dual EGF-like modules (Figure 8).252 The EGF-like modules between LG2/3 are proteolytically sensitive and cleaved by BMP1/Tolloid-like proteases to release LG3.461 The structure of LG3 has been solved463 and is an accessible biomarker464 for a vast (and growing) number of diseases,465−474 including breast cancer.475 The molecular processes of endorepellinmediated angiostasis and autophagy, and the emerging paradigm connecting both pathways will be examined.

Even though the direct interaction between lumican and MMP-14 catalytic domain was evidenced experimentally,406 no modeling study focusing on the formation of such a molecular complex has been reported. However, recent investigations focusing on the formation of MMP-14-catalyric domain (PDB 1BQQ), the lumcorin peptide417 and a set of lumican-derived peptides424 proposed an original molecular modeling and statistical methodology to characterize, at the atomic level, the interactions between peptides derived from lumican and MMP-14. Peptide/protein docking experiments were performed with the HEX software. Through their study, the authors evidenced two main interacting regions of MMP-14 catalytic domain: the catalytic site and the MT-loop. Peptides conformation targeting the catalytic site were statistically analyzed and revealed which MMP-14 residues are crucial in the complex formation and activation process. This type of in silico work should help scientists to propose, in the near future, the design of other peptides extracted from lumican and possibly other ECM PGs, reproducing the antitumor effect of lumican and lumcorin. The models do not integrate any posttranslational modifications (PTMs), such as N-glycosylated residues or the presence of glycosaminoglycan chains at the surface of the lumican protein. Grafting N-glycosylated chains on the surface of a SLRP, namely human lumican, highlights the steric hindrance generated by the glycosylated chains and illustrates how the distribution and orientations of the PTMs might impact the properties and interactions of lumican (Figure 7).

5. BASEMENT MEMBRANE PROTEOGLYCANS 5.1. Endorepellin: Dual Receptor Antagonism, Plethora of Functions

Proteoglycan diversity and functional variation comes in many forms; perhaps, best exemplified by the colossal basement membrane HSPGs.4,425 The genomic and promoter structures of perlecan HSPG2 gene are complex426,427 and unexpectedly X

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

5.2. Endorepellin Suppresses Angiogenesis as a Dual Receptor Antagonist

kinase activity as the tyrosine kinase inhibitor, SU5416, significantly abrogates autophagy.453 Autophagic induction is also independent of the α2β1 integrin as cells treated with LG3 do not exhibit measurable changes in established autophagic markers (Figure 8B).453 In terms of the autophagic gene expression induced by endorepellin, LG3 may actively suppress these targets453 thereby aligning closely with the antiautophagic properties of perlecan.454 Future work will dissect the molecular relationship of angiogenesis and autophagy. However, a biological analogy is emerging insofar as the perlecan parent molecule is proangiogenic and antiautophagic, whereas the converse of antiangiogenic and pro-autophagic properties holds true for endorepellin. Details surrounding this relationship have recently come to light,487 by demonstrating that endorepellin and Torin 1488 both inhibit vessel sprouting in aortic ring assays, an ex vivo model of angiogenesis, suggesting that endorepellin-evoked autophagy and similarly cleaved matrikines, such as the collagen type XVIII489 angiostatic fragment, endostatin,490 may inhibit tumor angiogenesis.487 Importantly, the angiostatic propensities of endorepellin are significantly attenuated by the AMPKα inhibitor, Compound C, suggesting an intersection of coordinate regulation between angiostasis and autophagy.

Endorepellin significantly and specifically impedes angiogenesis, capillary morphogenesis, and endothelial cell migration.447,476,477 Soluble endorepellin is a dual receptor antagonist by acting as a molecular tether for the simultaneous ligation of VEGFR2 and the α2β1 integrin (Figure 8).478 The molecular configuration of dual receptor antagonism positions the N-terminal LG1/2 module within IgG3−5 motifs of the VEGFR2 ectodomain, whereas LG3 binds the α2 I-domain of the α2β1 integrin (Figure 8).479−482 The mechanistic advance of “dual receptor antagonism” provides a molecular explanation for the exquisite sensitivity as subnanomolar concentrations of endorepellin elicit biological responses, and specificity of endorepellin on endothelial cells, the only cells that coexpress both receptors. The ternary endorepellin-VEGFR2-α2β1 integrin complex brings the SHP-1 tyrosine phosphatase, constitutively bound to the cytoplasmic tail of the α2 integrin subunit, in close juxtaposition with the unstructured VEGFR2 intracellular tails483 for rapid dephosphorylation (Figure 8A). VEGFR2 dephosphorylation results in heterotrimer complex inactivation, attenuation of downstream signaling effectors (Figure 8A)478 and caveolin-mediated internalization.484 Mechanistically, these events mirror those evoked by decorin (see above), suggesting convergent biological effects. Mechanistically, SHP1 mediated VEGFR2 dephosphorylation at Tyr1175485 ablates a key Shb docking site for subsequent PDK1/PI3K/Akt/ mTOR activation and PLC-γ nucleation,484 resulting in protracted angiogenic suppression (Figure 8A). The proangiogenic HIF-1α/VEGFA signaling axis is potently suppressed in a noncanonical, oxygen-independent manner,484 akin to decorin. Further, NFATc1 is inhibited downstream of PLC-γ, resulting in diminished Ca2+-dependent calcineurin activity and NFATc1 hyperphosphorylation, consistent with NFATc1 nuclear exclusion. The PKC/JNK/AP1 axis is also suppressed, culminating in decreased VEGFA expression and secretion (Figure 8A).478 Implicit within this model is a multiscale temporal dynamism lasting from minutes for receptor dephosphorylation to long-lasting transcriptional changes on the order of hours. Due to the inherently modular nature of endorepellin, its bioactivities can be physically separated by binding regions (e.g., LG1/2 vs LG3). The VEGFR2-binding domains LG1/2 are sufficient for attenuating VEGFA signaling,480 whereas LG3, the α2β1-binding region, is competent for actin cytoskeleton dissolution (Figure 8A).455,480

5.4. Endorepellin Evokes a Pro-Autophagic Gene Signature

Understanding the breadth of endorepellin-evoked autophagic gene regulation in endothelial cells is limited to individually determined PEG3, BECN1, and MAP1LC3A expression profiles (Figure 8B).453 A higher-resolution picture detailing the extent of autophagic gene regulation and the unique expression signature written by endorepellin was examined. Recent advances in autophagy indicate that stable and prolonged autophagic responses, such as those influenced by endorepellin, require sustained transcriptional output.346,491,492 NanoString, a next-generation digital PCR transcriptomics platform capable of “counting” individual, user-defined mRNAs, was used to evaluate a custom gene panel of 95 autophagy-related targets. Transcriptomic profiling revealed a 23-gene differentially expressed pro-autophagic gene signature evoked by endorepellin (Figure 6C).493 This study was focused on two coupregulated mitochondrial-associated genes, PARK2 and TCHP, encoding form Parkin and mitostatin proteins, respectively (Figure 8C). Coinduction of both proteins required the tyrosine kinase activity of VEGFR2 (Figure 8C). Considering the role of Parkin in the turnover of damaged mitochondria,366,494,495 it was found that relative to vehicle (Figure 9A), via the LG1/2 domains (Figure 9B), endorepellin (Figure 9C) is capable of mediating ΔΨm in a VEGFR2-dependent manner (Figure 8C). Depolarization was comparable with fluorophores FCCP (Figure 9D) or CCCP, established ΔΨm stimuli. JC-1 mitochondrial dye was used to assess ΔΨm. JC-1 accumulates within the inner mitochondrial membrane in response to ΔΨm. Upon loss of ΔΨm, JC-1 is monomeric and exhibits green fluorescence (Figure 9, panels B−D). Conversely, JC-1 aggregates proportionately to a high ΔΨm to form JC-1 aggregates that shift the JC-1 emission spectrum to a red fluorescence (Figure 8C, Figure 9A).496 Moreover, following ΔΨm, an interaction between mitostatin and Parkin as well as an increased association of mitostatin with an established Parkin receptor, mitofusin 2

5.3. Endorepellin Evokes Endothelial Cell Autophagy for Sustained Angiostasis

In a manner biologically analogous to decorin, nanomolar concentrations of endorepellin evoke bulbous, dually positive autophagosomes (Figure 6B) in endothelial cells similar to those formed after by canonical autophagic stimuli.486 Furthermore, endorepellin under nutrient-rich conditions, induces a stable and long-lasting pro-autophagic program converging upon the same autophagic core of VEGFR2, AMPKα, Peg3, Beclin 1, p62, and conversion of LC3-I to LC3II.453 As discovered for decorin, Peg3 acts as an essential mediator of the matrix-driven autophagic gene response. Refining our model of “dual receptor antagonism”, endorepellin acts as a partial VEGFR2 agonist, as endorepellin-evoked autophagy requires full VEGFR2 tyrosine Y

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

cytoplasmic domain (60−70% sequence identity). The ectodomain harbors two or three consecutive Ser-Gly sequences flanked by hydrophobic and acidic residues, which serve as attachment sites for HS and in some cases CS.501 Notably, the intact ectodomain can be proteolytically cleaved from the cell surface in the process of shedding (see section 6.3). The presence of the dimerization motif GxxxG in the transmembrane domain enables homoand hetero-oligomerization of syndecans,502 whereas the short cytoplasmic domains of all syndecans contain two conserved regions (C1 and C2) separated by a variable region (V) that is specific for individual syndecan members but conserved across species. The Cterminal EFYA amino acid motif present in all syndecans enables interactions with large PDZ-domain scaffolding proteins which link to the cytoskeleton.497 Apart from these common features, individual syndecan members have some unique properties, with syndecan-1 and -3 and syndecan-2 and -4 representing subfamilies. Syndecan-1 is expressed earliest during development and highly expressed on epithelia and plasma cells in adult vertebrates. Mesenchymal cells express syndecan-2, syndecan-3 is highly expressed by neuronal cells, and syndecan-4 is the most ubiquitously expressed member of the family.503,504 Of the five GAG attachment sites in the ectodomain of syndecan-1, three are located at the N-terminus and can be substituted with HS or CS, whereas two sites localize close to the membrane and are substituted with CS chains. GAG attachment site mutagenesis revealed that a reduction in the usage of syndecan-1 GAG attachment sites is associated with reduced cell-cell interactions and invasiveness.505 Furthermore, some ligand interactions, including binding of syndecan-1 to PT, MK, FGF-2, and lacritin appear to require either a defined cooperation (or absence, respectively) of either HS or CS.506,507 Moreover, degradation of HS by HPSE makes syndecan-1 more susceptible to cleavage by proteolytic enzymes, indicating a protective or regulatory role of the GAG chains for the core protein.508 While most of the ligand interactions of the syndecan-1 ectodomain are thought to be mediated by the GAG chains (as suggested by the comparably low conservation of the ectodomain AA sequence), at least two sequence motifs have been shown to be of particular relevance: functional data suggest that the hydrophobic amino acid sequence AVAAV, located 26 AA from the transmembrane domain, is involved in regulating cancer cell invasiveness,509 whereas a sequence covering AA 93−120 in human syndecan-1 termed “synstatin” mediates the lateral association with αν integrins and formation of a complex with IGFR (see Figure 11).510 In the syndecan-4 ectodomain, a conserved NXIP motif is required for cell adhesion properties.511 In all syndecans, the GxxxG sequence of the transmembrane domain mediates α-helical oligomerization, permitting formation of heterooligomers in a hierarchical manner.502 Apparently, homo-oligomer formation is less favored for syndecan-1 compared to the other syndecans, and heterooligomer formation with syndecan-1 is preferred by syndecan-2 and syndecan-3. A conserved phenylalanine residue in the transmembrane domain has been shown to be crucial for heterodimerization of syndecans.512 At the functional level, it is noteworthy that the transmembrane domain alone is capable of regulating focal adhesion disassembly, thus modulating cell migration,513 and Rac activity, cell migration, membrane localization of PKCα, and focal adhesion formation have been identified as readouts of hetero-oligomerization of syndecan-2 and -4.512,514 Notably,

Figure 9. Endorepellin evokes mitochondrial depolarization in endothelial cells. (A−D) Representative fluorescence micrographs depicting live-cell imaging of HUVEC (A) under vehicle conditions or following incubation with (B) LG1/2, (C) endorepellin, or (D) FCCP. HUVEC were cultured in nutrient-rich media and incubated with JC-1 to assess mitochondrial membrane potential. Note that endorepellin, LG1/2, and FCCP reduce mitochondrial membrane potential, as shown by JC-1 shift in fluorescence from red to green. Scale bar ∼20 μm.

was found (Figure 8C).249 Intriguingly, endorepellin increases the rate of autophagic flux as measured by the rate of LC3-II conversion in the presence of bafilomycin A1. Collectively, a unique autophagic expression signature that identified a novel role of endorepellin in regulating mitochondrial dynamics and, potentially, endothelial cell mitophagy has been determined. These new results synergize with the previously discovered in vivo data positing endorepellin as a potent antiangiogenic agent and could serve as the prototype for protein-based therapies in the future.

6. CELL-SURFACE PROTEOGLYCANS 6.1. Syndecans, Glypicans, and Transmembrane CSPGs: Primer to Cell-Surface PGs

Cell surface PGs are an important constituent of the glycocalyx and participate in cell−cell and cell−ECM interactions, enzyme activation and inhibition, and pleiotropic signaling processes, thereby regulating cell proliferation, adhesion, migration, and differentiation.17,497 The major classes of cell surface PGs comprise four members of the transmembrane HSPGs of the syndecan family, six members of the GPIanchored glypican family of HSPGs, and the CSPG CSPG4/ NG2 (Table 1). Additional cell surface PGs are so-called parttime PGs, as they can occur either in a non-GAG-substituted glycoprotein form or as a PG. The CSPG phosphacan, the TGF-β receptor, betaglycan, the VEGF coreceptor, neuropilin, and the HA receptor, CD44, serve as examples for this category.17 The structural organization of the most prominent members of cell surface PGs will be addressed in this section. Syndecans. The syndecans form a family of four type I singlepass transmembrane HSPGs, which are expressed by all adherent cells in a cell-type and developmental-specific pattern.498 As will be discussed in more detail in the following sections, syndecans act as receptors for the ECM, endocytic receptors, and most importantly as signaling coreceptors modulating RTK, chemokine, and morphogen signaling. In this way, they play an important role in regulating inflammation and angiogenesis, cell proliferation, differentiation, and cell adhesion and motility.497,499,500 Their core proteins are comprised of an ectodomain with comparably low amino acid sequence conservation (10−20% across the family) and a more highly conserved transmembrane region and Z

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

integrin and galectin binding sites and cleavage sites enabling proteolytic shedding.499 The large CSPG4 ectodomain interacts with numerous ECM components, integrins, and growth factors, including collagen type V and type VI, FGF and PDGF.17 While the unique transmembrane domain harbors a cysteine residue, the cytoplasmic domain is characterized by threonine acceptor sites for physphorylation by p44/42 MAPK and PKCα. Similar to the syndecans, the cytoplasmic domain of CSPG4 is capable of interacting with PDZ-domain proteins, including GRIP1, MUPP1, and syntenin.6 Betaglycan. Betaglycan (TGF-β type III receptor) acts as a coreceptor for TGF-β family cytokines.523 Interactions with activins, inhibins, GDFs, and BMPs define TGF-β-mediated functions of betaglycan in reproductive biology and epithelialto-mesenchymal transition. The single pass transmembrane PG contains several GAG attachment sites in its ectodomain, which also harbors a unique domain called zona pellucida-C, a transmembrane domain, and a short cytoplasmic domain with a high content of Ser/Thr residues, of which at least two are subject to phosphorylation.17 Similar to syndecans and CSPG4, the cytoplasmic domain of betaglycan contains a PDZ-domain binding sequence, allowing for potential interactions with scaffolding proteins such as β-arrestin. Finally, proteolytic cleavage of the ectodomain by granzyme B can convert betaglycan into a soluble effector molecule and is a means of downregulating TGF-β-mediated signaling.524 Phosphacan. Phosphacan is the soluble ectodomain of the single-pass type I membrane CSPG receptor-type protein tyrosine phosphatase β that plays a role in homotypic binding and heterotypic interactions with N-CAM in neurons, in neuronal-glia interactions and within stem cell niches.525 The ectodomain shows homology to α-carbonic anhydrase and contains fibronectin type III repeats and four consensus sequences for GAG attachment. The long transmembrane form of receptor-type protein tyrosine phosphatase β contains two tyrosine phosphatase domains in its cytoplasmic tail. CD44. The gene encoding the HA receptor, CD44, comprises 20 exons, of which 10 are subject to alternative splicing, thus generating several isoforms.24 While the most widely expressed standard isoform, CD44s, lacks a GAG chain, the CD44 HSPG contains the exon V3 which harbors GAG attachment sites. Apart from a substitution with GAG chains, the extracellular domain can be N- and O-glycosylated, which encourages its role as a part-time PG. CD44 mediates several functions of hyaluronan (see section 9) and exhibits coreceptor functions for EGF and hepatocyte growth factor (HGF) signaling.24,500 CD44 is a surface marker for several types of stem cells, including cancer stem cells.526 Notably, it has been shown that the HSPG variant of CD44 is implicated in nuclear targeting of the pluripotency-associated transcription factors Oct4, Sox2, and Nanog,527 providing a mechanistic link between CD44 V3 and the stem cell phenotype, with relevance to tumor progression and therapeutic resistance.24,526

the transmembrane domains of syndecan-1 and syndecan-3 contain a cleavage site for gamma-secretase, which has an impact on ectodomain shedding and signal transduction (see Figure 12).515,516 It is the most highly conserved regions of the syndecan core proteins, the cytoplasmic domains, where the largest structural and functional divergence between the syndecan family members is observed. Several conserved tyrosine residues have been identified in the cytoplasmic domains, which are known to be phosphorylated. The C1 domain is rich in basic amino acids and contains the juxtamembrane sequence RMKKK, which acts as a signal for tubulin-dependent nuclear localization517 and represents a consensus sequence for proteolytic cleavage mediated by γsecretase.515 It also contains a conserved serine residue which can be phosphorylated. Binding interactions with tubulin, ezrin, cortactin, Fyn, and Src have been demonstrated for the C1 domain.497 The variable V region separating the conserved C1 and C2 region of the syndecan cytoplasmic domains defines specific functions of the individual syndecans. This has been most intensely studied in the case of the syndecan-4 Vregion, which interacts with α-actinin, PIP2, and PKCα, thereby regulating focal adhesion assembly and disassembly and promoting syndecan oligomerization.500 Syndecan-1/HSdependent activation of cytosolic PKC in fibroblasts results in phosphorylation of serine 714 in transient receptor potential canonical channel-7 (TRPC-7), controlling cytosolic calcium equilibria, and subsequently cytoskeletal organization and myofibroblast phenotype.518 In the case of syndecan-1, the V-region plays an important role for the actin-bundling protein fascin and cell spreading,519 whereas the V region of syndecan2 serves as a substrate for PKCα and protein kinase A. Apart from mediating interactions of its EFYA PDZ-domain binding motif with the scaffolding proteins syntenin and CASK, the C2-domain binds to TIAM1 and synectin.500 Glypicans. The glypicans are a family of six GPI-anchored cell surface HSPG, which are characterized by an ∼50 kDa domain that contains 14 highly conserved cysteine residues and a region near the plasma membrane that contains 3−5 Ser-Gly GAG attachment sequences.497,520 By presenting GAG chainbound morphogens and growth factors to their cognate receptors, glypicans modulate numerous signaling processes with relevance to tumor cell proliferation and angiogenesis.17 The membrane-proximal location of the GAG chains, the globular (vs extended) structure of the core proteins, and the lipid-mediated membrane attachment distinguishes the glypicans from syndecans at the functional level.497,520 As will be discussed in the next sections, the membrane-proximal location of the glypican GAG chains allows for functional interactions with cytokines and morphogens such as Hh, Wnt, and FGFs.520,521 Similar to syndecans, the ectodomain of glypicans can be shed from the plasma membrane, however, via distinct mechanisms (see section 6.3).499 CSPG4/NG2. The human melanoma-associated CSPG4 encodes a 2322 AA long single pass type I transmembrane protein to which CS GAG chain is attached.17,522 The term NG2 refers to the rat ortholog, nerve/glial antigen 2. CSPG4 and NG2 play important roles in tumor angiogenesis, in oligodendrocyte precursor and cancer cell migration.6,17 CSPG4 has a substantial ectodomain, which can be subdivided into an N-terminal domain (D1 subdomain) containing laminin-like globular repeats, a central subdomain (D2) containing 15 cadherin-like domains, and a carbohydratesubstituted juxtamembrane subdomain (D3) containing

6.2. Cell-Surface PGs as Pleiotropic Integrators of Signaling Processes

As pointed out in section 3.1, the GAG chains of PGs interact with a plethora of growth factors, chemokines, and morphogens. This is of particular functional importance in the case of cell surface PG, where the GAG-bearing extracellular domains of PGs can act as so-called coreceptors.500 A prototype example for this function are the AA

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 10. Cell surface HSPGs as classical coreceptors for growth factor and chemokine signaling.497,530,531 (A) Syndecans act as coreceptors for FGF-mediated signaling. In the absence of ligand, the FGF receptor (FGFR) is a nonsignaling monomer. The HS chains of cells surface PGs, such as syndecans enrich the ligand, FGF, at the cell surface in a HS-sulfation-specific manner. The HS chain, FGF, and FGFR form a ternary complex that facilitates FGFR dimerization and subsequent autophosphorylation of the receptor at its cytoplasmic domain, which exhibits tyrosine kinase activity. The newly generated tyrosine residues provide docking sites for Src homology domain containing adaptor proteins, such as GRB2, which trigger a series of serine/threonine phosphorylation events in the MAPK cascade. One result of the signaling cascade is activation of transcription factors, which induce the expression of various gene products with relevance to cell proliferation, migration, invasion, and angiogenesis. Example shows the role of HSPG in a cis conformation (on the same cell as the RTK), but activation in trans in a noncell-autonomous manner is possible. (B) Role of cell surface HSPGs in chemokine-mediated leukocyte recruitment. Endothelial HSPGs such as syndecans are capable of forming chemokine gradients by enriching these chemotactic cytokines at the cell surface. In this way, the chemokine is efficiently presented to chemokine receptors at the surface of leukocytes. Activation of these G-protein-coupled heptahelical receptors (GPCRs) can trigger various signaling events (e.g., JAK/Stat, Ras, ERK, and Akt). Here, the activation of a G-protein is shown, which results in activation of GPCR-dependent kinases which trigger chemotactic migration of the leukocyte, allowing for transmigration through the endothelium (diapedesis).

syndecans and their modulation of RTK-signaling; via their HS chains, syndecans are capable of forming a ternary complex with the ligand (e.g., FGF2) and the RTK (e.g., FGFR).498,528 In this context, the HS chain of the PG immobilizes the ligand and increases its local concentration, is capable of changing its conformation, and presents it to the RTK, thus resulting in increased receptor dimerization and stimulation of intracellular signaling (Figure 10A).497,529 This coreceptor function for growth factors has been demonstrated for HSPGs (syndecans, glypicans), CSPG4, and the HA receptor CD44.24,497,520 HSPGs also modulate signaling via a different class of receptors, GPCRs,530 as exemplified by chemokine signaling (Figure 10B). In this case, cell surface HSPGs not only assist in forming chemokine gradients at the endothelial surface, which serve to guide leukocytes from the circulation to sites of inflammation via binding of positively charged amino acids to the GAG530 but also are thought to facilitate signaling through the GPCRs by promoting multimerization of the ligand.532,533 In the case of CXCL12, HS was shown to present different chemokine isoforms to the cognate receptor CXCR4 via distinct domains,534 whereas analysis of the solution structure of CXCL13 revealed that the chemokine can be presented to its receptor in a HS-bound form, and that the association with HS occurs via two clusters located in the α-helix.532

The classical concept of GAG-mediated coreceptor functions of cell surface PGs has recently been extended to other areas of signal transduction. An area of intense research is the modulation of integrin signaling by cell surface PGs. Integrins are a class of heterodimeric transmembrane receptors which provide a link between the ECM and the intracellular cytoskeleton.535 Similar to HSPGs of the syndecan family, their short cytoplasmic domains are not capable of direct signaling; however, they can act as docking sites for the assembly of large signaling complexes.535 Increased activation of focal adhesion kinase (FAK), a downstream effector of integrin function, has been observed in syndecan-1-depleted MDA-MB-231 breast cancer cells, resulting in increased resistance to radiotherapy.536 Moreover, syndecan-1-deficient mice exhibit increased inflammation in a variety of experimental models (contact allergy, colitis, myocardial infarction, anti-GBM glomerulonephritis) apparently due to a more efficient interaction of leukocyte integrins with their endothelial receptors in the absence of the PG.531,537 These data suggest overlapping functions and a mechanistic link between syndecans and integrins. Indeed, the ectodomain of syndecan-1 contains a binding site for αν integrins,538 which is the molecular basis for the regulatory impact of syndecan-1 clustering on ανβ3 and ανβ5 integrins in a variety of cell types.500,510 Association of the IGF1-receptor with the syndecan-1-αV-integrin complex plays an important mechaAB

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

integrin by the EGF receptor typrosine kinase HER2 depends on interaction of the cytoplasmic domain of the β4 integrin subunit with syndecan-1, whereas its activation via HER1/ EGFR depends on syndecan-4, differentially affecting cell motility and survival.40 The syndecan-1 ectodomain is also capable of modulating the interaction with the ECM by interacting with αν and α5β1 integrins,500 whereas Srcmediated phosphorylation of the syndecan-4 cytoplasmic domain is a mode of regulating integrin recycling.540 A different mode of integrin association has been described for syndecan-2, which indirectly associates with β1 integrins via association with CD148, an integral membrane tyrosine phosphatase.541 Notably, syndecans are not the only family of transmembrane PGs capable of modulating integrin function. For example, the HSPG variant of CD44 can interact with leukocyte β2 integrin via its HS chains,542 whereas phosphorylation of NG2/CSPG4 at distinct threonine residues determines the colocalization and cellular distribution of β1integrins, thus directing their specific downstream effects on cell proliferation and migration.49 NG2/CSPG4 is capable of activating β1 integrins both in a cis and a trans conformation.543 Apart from a modulation of growth factor-, chemokine-, and integrin-mediated signaling, cell surface PGs are known to regulate developmental processes and stem cell function as part of the signaling mechanism of morphogens, specifically Wnt and Hh families that directly act on cells to induce distinct cellular responses in a concentration-dependent manner.497,544 In addition, a modulation of the stemness-related notch signaling pathway has emerged as a novel function of cell surface PGs. Data from model organisms have demonstrated that cell surface PGs play an activating role for Wnt, which signals through the Frizzled family of seven pass transmembrane receptors.545 Canonical Wnt signaling involves the coreceptor LRP5 and cytoplasmic accumulation of β-catenin and its nuclear translocation, resulting in an activation of transcription factors of the TCF/LEF family. In contrast, the noncanonical Wnt signaling pathways do not involve β-catenin and can be subdivided into the planar cell polarity pathway, regulating cell polarity and migration via cytoskeletonmodulating Rho-GTPases and the Wnt/calcium pathway, which utilizes activation of the PLC/DAG/IP3/Ca2+-pathway to regulate cell migration and proliferation.546 Cell surface HSPGs of the glypican and syndecan family are linked to altered Wnt signaling, exhibiting an impact on embryonic development and malignant disease. In Drosophila, several mutants in HS biosynthetic enzymes (e.g., sugarless, sulfateless) and in the Drosophila homologue of glypican, dally, exhibit signaling defects of wingless, the fly homologue of mammalian Wnt,497 whereas data in the frog Xenopus showed that the secreted Wnt modulator R-spondin-3 binds to syndecan-4 and induces clathrin-mediated endocytosis of Wnt5a-receptor complexes, leading to their activation.547 With regard to canonical Wnt/β-catenin-signaling, it has been demonstrated that syndecan-4 and its ECM substrate fibronectin inhibited Wnt/β-catenin at the cell membrane through regulation of LRP6, suggesting differential effects of this HSPG on canonical and non-canonical Wnt pathways.548 Moreover, syndecan-4 coordinates Wnt/JNK and BMP signaling to regulate Xenopus foregut progenitor development, simultaneosuly influencing signaling through seven pass membrane receptors and fibronectin-dependent TGF-β-related pathways.549 Recent results in mammalian cells suggest that

nistic role in this process, as it leads to autophosphorylation of the RTK and subsequent integrin activation (Figure 11).510,538 Inhibitory peptides directed against the syndecan-1-αν integrin binding site called synstatins have shown promising therapeutic results in preclinical models of breast cancer and myeloma.510,538 Further examples for a modulation of integrin signaling by syndecans are the finding that activation of α6β4

Figure 11. Role of syndecans in integrin activation.500,510,531,536−539 Syndecans bind to similar ligands in the ECM as integrins (e.g., fibronectin, laminins), albeit to different domains (integrin- or cell binding domains vs Hep binding domains).500 Moreover, they modulate similar downstream signaling pathways (e.g., Rho-GTPase and FAK-mediated pathways)500,536,539 and both act as matrix adhesion receptors and mediators of cell motility. Shown here is the example of ανβ3/5 integrin activation by syndecan-1, which involves an activation of the IGF-I.510,538 The upper panel shows the components of the signaling complex in its inactive state, whereas the bottom panel shows the assembled signaling unit. Syndecan-1 acts as a classical coreceptor in activation of the IGFR by its cytokine ligand IGF. IGFR activation results in an inside-out-activation of the integrin, which acquires an activated state allowing for engagement of large matrix glycoproteins (e.g., vitronectin, fibronectin, fibrinogen, and laminins), which also interact with the HS chains of syndecan-1. A particular feature of ανβ3/5 integrin is its binding to a region in the extracellular domain of syndecan-1 that is homologous to the peptide drug synstatin, allowing for lateral association with the HSPG. This process allows for activation of the actin cytoskeleton and signaling events that promote cell proliferation, invasive growth of tumor cells, and cell survival.538 Synstatin is an antiangiogenic and antitumoral drug which can inactivate this signaling complex via competitive binding. Apart from this mode of signaling, other modes of integrinsyndecan-interactions have been described, including interactions of the cytoplasmic domains of syndecan-1 and syndecan-4 with α6β4integrins,40 an increased activity of FAK and β1-integrins in breast cancer cells depleted of syndecan-1,536,539 and increased HSdependent interactions of leukocyte integrins (β2) with their ligand ICAM-1 in the absence of syndecan-1.531,537 AC

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 12. Cell surface HSPGs as modulators of stemness-related signaling pathways (see refs 520, 526, 557, 558, 560, and 561). (A and B) Differential role of glypican-3 and glypican-5 in Hh signaling. In the absence of stimulating Hh ligand, the GPCR Smoothened is repressed in a mechanism involving the membrane protein Patched. Upon binding of Hh to Patched, Smoothened is no longer repressed and able to signal. Hh binds to glypicans, and its release from glypicans may depend on sheddases.557 Glypican-5 binds Hh via its HS chains and presents it to Patched, thus exhihiting classical HSPG coreceptor function (A). In contrast, glypican-3 competes with Patched for Hh binding, thus downregulating signaling through this pathway (B). This pathway is modulated by convertases in the case of glypican-3, and by the sheddase notum. (C) The role of syndecans in Notch signaling. Activation of Notch occurs upon juxtrakrine binding of opposing ligands (delta, serrate) on neighboring cells. This activation triggers sequential cleavage by the protease TACE (TNF-α converting enzyme) at the Notch ectodomain and intramembranous cleavage by γ-secretase. These cleavage events result in the cytoplasmatic release of the Notch intracellular domain (NICD), which can associate with accessory proteins to form a transcriptional regulator which activates stemness-related transcriptional programs. Although the mode of action of HSPG in this process is not fully clear, an involvement for 3-O-sulfated HS of syndecan-3 in Notch signaling in the muscle and of syndecan-1 and multiple Notch members in breast cancer have been documented.526,560,561 Notably, the same sheddases that act on Notch and activate this pathway do also cleave syndecans, providing a means of modulating the signaling pathway via downregulation of the HSPG coreceptor.499,500

noncanonical Wnt signaling induces chondrocyte dedifferentiation through Frizzled-6 and a DVL-2/B-Raf/CaMKIIα/ syndecan-4 axis, providing a possible mechanistic explanation of the involvement of Wnt signaling in osteoarthritis.550 Overall, these results provide strong evidence for involvement of syndecan-4 in various Wnt-related pathways. A role for syndecan-1 in Wnt-dependent cancer stem cell function has been revealed in studies on syndecan-1-deficient mice and human breast cancer cells. Pioneering data from the group of Caroline Alexander had demonstrated that syndecan1-deficient mice exhibit a reduced Wnt1-responsive precursor cell pool in the mammary gland, which was the underlying mechanistic cause for the resistance of juvenile syndecan-1deficient mice to experimentally induced breast cancer.551,552 Data from syndecan-1 siRNA-depleted human breast cancer cells demonstrated that syndecan-1 deficiency reduced the cancer stem cell population and formation of mammospheres in vitro via regulation of LRP-6 and IL-6-mediated STAT3 signaling.526,553 Data in CRISPR/Cas9-mediated knockout of the HS biosynthetic enzyme EXT1 in multiple myeloma cells suggest that syndecan-1 plays a classical coreceptor role in Wnt signaling, as this HSPG promoted the presentation of Wnts and R-spondins in a GAG-dependent manner.554 Another developmentally relevant signaling pathway is mediated by the stemness-related Hh proteins, which act as classical diffusible morphogens. Drosophila Hh and its mammalian homologues Sonic (Shh), Indian (Ihh), and Desert hedgehog (Dhh) are secreted HS-binding morphogens

that bind to the 12-pass transmembrane receptor Patched and a coreceptor, which depresses the GPCR Smoothened, controlling a variety of developmental processes through activation of Gli family transcription factors. A noteworthy structural feature of Hh is its lipidation, as the C-terminus of the mature protein is covalently attached to cholesterol, while its N-terminus harbors a palmitic acid residue.555 Similar to the Wnt pathway, initial results on a role of HSPG in this signaling pathway were generated in Drosophila mutants of HS biosynthesis, including the HS polymerases EXT1 and EXT2. Dysregulated expression of downstream signaling components of the Hh pathway in the uteri of mice deficient in the isoforms of the HS biosynthetic enzyme NDST.556 While Gli1 expression is affected by syndecan-1 knockdown in human breast cancer cells,526 the glypican family of HSPGs appears to be responsible for most HS-related effects on Hh signaling.520 Indeed, the HS chains of glypicans play a mechanistic role in stimulating Shh processing and release.557 Notably, different members of the glypican family have enhancing or inhibitory effects on Hh signaling (Figure 12). For example, glypican-5 stimulates rhabdomyosarcoma proliferation by increasing of Shh binding to Patched (interacting with both partners in a GAG-dependent manner), whereas glypican-3 does not bind Patched but competes with the receptor for Hh binding via a mechanism that depends on processing by furin-like convertases.520,558 Additional glypican members that modulate Hh signaling are glypican-1, which acts as a coreceptor for Shh in repulsive guidance of commissural neurons,520 and glypicanAD

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

6, which promotes the growth of developing long bones by stimulating the Hh pathway.559 Overall, these data provide not only strong evidence for a role of glypicans in Hh signaling but also demonstrate the context-dependent functions of different members of this lipid-anchored family of cell surface HSPGs. The final example for a modulation of stemness-related pathways by HSPGs is represented by the Notch family of transmembrane receptors (Figure 12C). Notch receptors are a family of four transmembrane proteins which interact with their membrane-bound ligands of the delta-like and jagged families in a juxtacrine manner, resulting in two consecutive cleavage events by ADAM10/17 at the extracellular juxtamembrane domain and by γ-secretase in the transmembrane region. These proteolytic events lead to a release of the Notch intracellular domain which associates with additional proteins that migrate into the nucleus where they activate transcriptional programs involved in mediating stemness-related functions.526,562 Initial results obtained in Drosophila pointed at a role of HS in Notch signaling, as this pathway was impaired in mutants of HS 3-O-sulfotransferases.560 Apparently, syndecans are functionally linked to this pathway. The expression of syndecan-2 is induced in smooth muscle cells when these cells contact endothelial cells expressing Notch-2 and Notch-3, or with the ligands themselves. In turn, notch activity is attenuated in smooth muscle cells deficient in syndecan-2, suggesting reciprocal regulation of syndecan-2 and Notch signaling in this system.563 Experiments in syndecan-3-deficient satellite cells demonstrated that syndecan-3 interacts with notch and is required for its proteolytic processing, suggesting that the two transmembrane proteins cooperate in regulating the satellite cell pool and myofiber size.561 Finally, syndecan-1 has been linked to Notch signaling in the context of malignant disease. Data on inflammatory breast cancer tissues and cell lines demonstrated that syndecan-1 expression correlated with Notch-1, Notch-4, and CD44 expression in patient tissues, and that siRNAmediated knockdown of syndecan-1 in inflammatory breast cancer cells resulted in downregulation of all notch receptors and of the downstream transcription factor Hey-1. Pharmacological inhibition of Notch activation resulted in a phenocopy of the effect of syndecan-1 depletion on inflammatory cytokine and Hey-1 expression suggesting a functional interplay of these signaling pathways.526 Indeed, Notch activation by δ-likeligand-1 promotes the proliferation multiple myeloma cells predominantly in the syndecan-1 (CD138) positive subpopulation.564 While these data provide a clear link between the notch signaling pathway and syndecans and mark it as a therapeutic target, the detailed molecular mechanisms are still elusive.

Shedding of a membrane PG is therefore a means of eliminating the PG coreceptor function for various signaling processes at the cell surface, a mechanism for modulating cell adhesion and cell-matrix interactions, and in case of a paracrine function of the cleaved ectodomain, a way of either neutralizing signaling molecules in the vicinity of the cell, or of promoting signaling processes in a neighboring cell if the PG ectodomain acts in a trans configuration.497,499,500 Therefore, shedding can have important consequences for cell behavior. For example, shedding of syndecan-1 has been suggested to mark a switch from a proliferative to an invasive state of breast cancer cells,565 the soluble syndecan-1 ectodomain enhances proteolysis and delays skin wound repair in mouse models,566 and syndecan-1 shedding is exploited by microbes as part of the infection mechanism, as it promotes escape from the immune system and modulates chemokine function.567 Ectodomain shedding of transmembrane PGs has been described for syndecan-1, -2, and -4, for NG2/CSPG4, phosphacan, and for betaglycan.497,499,500 In the case of transmembrane PGs, cleavage of the intact ectodomains usually occurs at sites close to the plasma membrane and is mediated by proteases of the ADAM, ADAMTS, matrilysin, collagenases, gelatinases, and the serine proteases plasmin and thrombin.567 Shedding can be constitutive or induced by a wide range of stimuli which involve signaling through the PKC, tyrosine kinase, and the MAP kinase pathways.497,499,500 A modulation of morphogen signaling through shedding has been described for glypicans; however, in contrast to the aforementioned transmembrane PGs, shedding of these lipidanchored PGs requires a different mechanism. Cleavage of the GPI anchor by the phospholipase Notum has been described for glypican-3, -5, and -6568 and negatively affects morphogen gradient formation and Wnt signaling in Drosophila, whereas Hh signaling is apparently activated via induction of endocytic internalization of glypican−ligand−receptor complexes.569,570 An interesting question concerns the fate of the transmembrane and cytoplasmic regions of the shed PGs. In the case of syndecan-1, it has been shown that the transmembrane region is cleaved by gamma-secretase and ultimately processed by the proteasome.516 Interestingly, the cleavage process is still of functional importance and not a mere act of degradation of a nonfunctional protein remnant, as the PG fragment inhibits cell migration and invasion of lung cancer cells by activating multiple intracellular kinases, including Src kinase, FAK, and Rho-GTPase. While shedding of cell surface PGs is a way of modulating their function to generate a wide range of additional effects, nuclear localization of PGs has been discussed as a way of regulating gene expression. Indeed, several membrane PGs contain amino acid motifs that represent nuclear localization sequences (NLS).571 In the case of syndecan-1, the conserved juxtamembrane sequence RMKKK represents a NLS, in which the arginine residue plays a particularly important role for tubulin-dependent internalization of the HSPG.517 Several signaling pathways have been shown to regulate nuclear localization of syndecan-1, including PKC, TGF-β, and possibly FGF2-triggered dephosphorylation of phosphorylation sites within the HSPGs cytoplasmic domain.571 Apart from the syndecans, CD44 (and its PG variant) harbors a NLS (RRRCGQKKK) in its cytoplasmic tail. CD44 reaches the nucleus via nuclear pore complex after the protein is endocytosed and sorted in endosomes. While a potential NLS [KRRR(G/A)] is also present in glypicans, it has been

6.3. From Ectodomain Shedding to PDZ-Domain Interactions: Molecular Modes of Regulating Cell Behavior

In the previous sections, the role of cell surface PG as multifunctional modulators of signaling processes has been highlighted. The mechanistically interesting features in more detail will be addressed. One mechanism, which is of particular functional relevance is ectodomain shedding, the cleavage of intact extracellular domains of cell surface PGs by proteases and, in the case of glypicans, phosphoslipases. This mechanism converts the membrane bound PG ectodomain into a soluble effector molecule which is capable of binding to the same extracellular ligands and which can function as a paracrine modulator of signaling and cell-ECM-interactions.497,499,500 AE

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

behavior, for example, a switch from exosome formation and syndecan recycling to changes in cell adhesion and motility.

discussed that the site may be blocked by the HS chains which are located in close proximity.571 Therefore, degradation of the GAG chain may be a requirement for enabling nuclear trafficking of glypicans. Indeed, in the case of syndecans, it has been shown that the HS-degrading enzyme HPSE regulates nuclear localization of the PG, albeit in a negative manner. This is of functional relevance, as a reduction of nuclear syndecan-1 enhances histone acetyltransferase activity, and this epigenetic mode of regulation results in an increased expression of genes which drive an aggressive tumor phenotype, including the protease MMP-9 and the cytokines VEGF and HGF.572 While it is still a matter of debate if the GAG chains, or only the protein moieties of PGs reach the nucleus, it is easily conceivable by nature of the high negative charge of both GAGs and DNA that nuclear GAGs may bind to several positively charged proteins that regulate the accessibility and the functional status of DNA. While it is widely accepted that the integrin family of heterodimeric transmembrane proteins provides a link between the ECM and the intracellular cytoskeleton, it is noteworthy that all transmembrane-anchored PGs are capable of interacting with large submembranous cytoplasmic scaffolding proteins via their PDZ-domains, thus exhibiting a similar function.500 PDZ domain proteins contain an approximately 90 AA motif with homology to the postsynaptic protein PSD-95, the junctional drosophila protein Discs-large, and the tight junction protein zonula occludes ZO-1. The cytoplasmic domains of several transmembrane PGs bind to the class I PDZ domain of synectin (GIPC/semcap-1), including syndecan-2, syndecan-4, betaglycan, and neuropilins.500 It has been discussed that this interaction may control cell adhesion and matrix assembly in cooperation with α5β1 integrins and may also be involved in syndecan-2-dependent recycling of betaglycan.500 Other PDZ-domain proteins interacting with transmembrane PGs include the syndecan-2 binding partner synbindin, which is involved in the regulation of postsynaptic vesicle trafficking and the partial regulation and activation of the MAPK pathway in gastric cancer and NG2/CSPG4, which has been discussed to be clustered via interactions with the PDZ-domain protein GRIP (glutamate receptor interaction protein).573 Moreover, binding of CD44 to its extracellular ligand HA allows for interactions with the type II PDZ motif of LARG (leukemia-associated RhoGEF), presumably leading to receptor clustering, which in turn has an impact on RhoGTPase signaling to the cytoskeleton.574 Interactions of cell surface PGs with the PDZ-domain protein syntenin have recently become a field of extensive study due to their relevance to exosome formation (see section 9). Syntenin is characterized by the presence of a tandem PDZ structure, which interacts with the EFYA motif of all syndecans, and the QYWV sequence of NG2/CSPG4.575,576 While syntenin is present in cell adhesion sites, it also participates in the recycling of syndecans via a mechanism that involves the endosomal GTPase ARF6 and differential interactions of the PDZ1 and PDZ2 domains with the EFYA motif of syndecans and the lipid phosphatidylinositol 4,5bisphosphate.577 Notably, the binding of transmembrane PGs to PDZ-domain proteins is subject to regulation by signaling cascades, as exemplified by syndecan-4: when its cytoplasmic serine residue is phosphorylated, the cytoplasmic domain undergoes a conformational change which does no longer allow for interactions with syntenin, while interactions with synetcin are still possible,578 thus determining changes in cell

6.4. Cell-Surface PGs as Diagnostic Markers and Therapeutic Targets in Malignant, Inflammatory, and Neurodegenerative Diseases

Due to their potent role to regulate multiple cell functions including adhesion, migration, and signaling, cell surface PGs have emerged as potent diagnostic and prognostic markers in several diseases. Furthermore, they appear as possible therapeutic targets for disease treatment by using several approaches including immunotherapy, use of GAG mimetic compounds and antagonists, novel binding peptides, and GAG biosynthesis competing xylosides. 6.4.1. CSPG4: Prognostic Potential and Target for Immunotherapy in Malignancies. CSPG4 is considered a valuable prognostic factor associated with disease progression in several malignancies including melanoma, sarcomas, and chordoma, malignant mesothelioma, glioblastoma, acute myeloid and lymphoblastic leukemia, breast, pancreatic, and oral, head, and neck squamous cell carcinoma.579−590 The elevated expression of CSPG4 by tumor cells is correlated with shorter time of disease-free and overall survival as well as with development of metastasis.584 CSPG4 is included in the list of cancer antigens that have been evaluated for the development of therapeutically effective cancer vaccines.591 Direct immunotargeting of CSPG4 with the anti-HMW-MAA monoclonal antibody (mAb) 225.28S inhibits human melanoma tumor growth in SCID mice and has been suggested as a reagent to be applied for passive immunotherapy to patients with malignant melanoma.592 An antibody against CSPG4 fused to soluble human TRAIL has also been shown to control melanoma growth due to combined inhibition of tumorpromoting CSPG4-signaling and activation of apoptosis.593 Similarly, a CSPG4-specific single-chain antibody fragment fused to a functionally enhanced form of the microtubuleassociated protein (MAP) tau has been tested against triple negative breast cancer cells displaying a significant tumor regression in vivo.594 Immunotoxins that specifically bind CSPG4 eradicate tumor cells by inducing apoptosis and inhibiting tumor growth.595,596 The CSPG4-targeting immunotoxin 225.28-saporin can be combined with photochemical internalization to create a specific and light-controlled treatment approach.596 Another strategy is the production of chimeric antibodies fused to superantigens such as staphylococcal enterotoxin A to activate T-cells. Treatment with the chimeric antibody induced HMW-MAA-specific tumor growth reduction by redirecting T-cells cytotoxicity to melanoma cells.597 Bispecific antibodies that bind CSPG4 and receptors mediating T-cell activation, such as the TCR/CD3 complex and costimulatory CD28, redirect T-cells cytotoxicity toward tumor cells.598,599 T-cells genetically engineered with CSPG4specific chimeric antigen receptor (CAR) also encoding the CD28 costimulatory endodomain demonstrate significant antitumor activity against a variety of solid tumors ex vivo and in vivo emerging CAR-CSPG4 T-cell immunotherapy as an effective treatment modality.598,600 6.4.2. Complex Roles of Glypicans in Diseases and Their Targeting. Prognostic Potential of Glypicans in Malignancies. Glypicans are expressed in a cell- and contextdependent manner and play a dual role either fostering or suppressing tumorigenesis.18,520 Elevated expression of glypican-1 is evident in various tumors including breast, pancreatic, AF

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

3-derived peptide vaccine therapies have been developed, and clinical trials have been performed in patients with ovarian clear cell carcinoma, hepatocellular carcinoma, and pediatric solid tumors with promising outcomes.628,646−648 Glypicans are Implicated in Neurodegenerative Diseases. It has been shown that glypican-1 is codeposited with βamyloid peptide in neuritic plaques in Alzheimer’s disease. HS chains of glypican-1 not only bind to fibrillar β-amyloid protein to evoke its deposition but also accelerate neuronal cell death.649−651 Glypican-5 polymorphisms are also associated with inflammatory demyelinating diseases such as multiple sclerosis and neuromyelitis optica.652−654 6.4.3. Syndecans: Potent Regulators of Malignancies, Inflammatory, and Neurodegenerative Diseases. Syndecans as Prognostic Indicators and Therapeutic Targets in Malignancies. Conflicting data on the correlation of syndecan1 on tumor cell surface with tumorigenesis have been described in malignancies such as breast, nasopharyngeal, colorectal, endometrial, and prostate cancer.6,20,655 In certain tumors including mesothelioma, basal and squamous cell, lung, head and neck, gastric, hepatocellular, cervical, and bladder carcinoma, the presence of syndecan-1 on the cell surface of tumor cells has been correlated with favorable prognosis.6,20,655 The loss of cell surface associated syndecan-1 evokes tumorigenesis in these tumor types. On the other hand, the elevated expression of syndecan-1 by tumor cells promotes disease in pancreatic, ovarian, and thyroid cancer, liposarcomas, multiple myeloma, and lymphomas.6,20,655 A strong positive correlation has been demonstrated between the elevated levels of soluble syndecan-1 and the expression of sydecan-1 by stromal cells with the progression of malignancies and unfavorable prognosis.6,655−659 The expression of syndecan-2 increases in various tumors including melanoma, lung, ovarian, colon, prostate, esophageal squamous cell carcinoma, and osteosarcoma.6,503 Higher expression of syndecan-2 is an unfavorable prognostic indicator for disease progression in prostate660,661 and esophageal squamous cell carcinoma.662 Syndecan-4 is upregulated in breast504,663 and colon cancer,664 osteosarcoma,665 melanoma,666 glioblastoma,667 testicular germ cell tumors,668 and malignant T-cells.669 Increased stromal staining of syndecan-4 in seminomas and reduced levels in tumor cells in nonseminomatous germ cell tumors are related to increased metastatic potential.668 In contrast, the elevated expression of syndecan-4 is associated with metastatic disease in osteosarcoma665 and shorter survival in glioblastoma.667 Single mAbs or drug-conjugated antibodies targeting syndecan-1 have been successfully used alone or as combined therapy with other agents to eliminate melanoma,670 triple negative breast cancer,671 multiple myeloma,672−675 in in vitro and in vivo studies. Vaccination with multiple HLA-A2-specific peptides of XBP1, syndecan-1, and SLAMF7 highly expressed in multiple myeloma induced antigen-specific cytotoxic T-cells with activity against myeloma cells in an HLA-A24 restricted manner and serves as a promising approach to target malignancies expressing these antigens.676,677 Genetically modified NK cells expressing CAR consisting of an antisyndecan-1 single chain variable fragment fused to the CD3ζ chain have been also developed and applied against multiple myeloma cells in vitro and in xenograft mouse model effectively.678

colorectal, uterine cervical, and esophageal squamous cell carcinoma and glioblastoma.6,41,601−608 Glypican-1 is a poor prognostic indicator associated with increased chemoresistance in patients with esophageal squamous cell carcinoma, pancreatic cancer, and glioblastoma.602,605,609 The levels of circulating glypican-1 enriched exosomes has been shown to be associated with severe clinical status and poor prognosis in colorectal carcinoma and pancreatic cancer.603,604,607,610 Methodologies using electrokinetic microarray and nanoparticle based chips have been developed to measure glypican-1 positive exosomes in liquid biopsies to improve early stage cancer diagnostics.607,611,612 Glypican-2 has been identified as an oncoprotein elevated in neuroblastoma that is correlated with poor overall survival.613,614 Glypican-3 exerts diverse biological roles in tumorigenesis. Decreased expression of glypican-3 related to progression of several malignancies including breast, gastric, renal cancer, and malignant mesothelioma.43,615−618 Loss of glypican-3 is associated with poor overall survival in gastric cancer, suggesting a tumorsuppressor role for glypican-3.43 On the other hand, the expression of glypican-3 is markedly increased in numerous malignancies including Wilms tumors,619 melanoma,620 thyroid,621 lung,622,623 hepatocellular,42,624−627 and ovarian clear cell carcinoma.628 Glypican-3 is suggested as a useful marker for early diagnosis of melanoma.620 High expression of glypican-3 is related to lower differentiation of cancer cells and the presence of lymph node metastases in lung cancer.622,623 Similarly, the elevated levels of glypican-3 are poor prognostic indicator for disease progression in hepatocellular carcinoma.42,625−627,629 Elevated expression of glypican-5 evokes tumor progression and metastasis in salivary adenoid cystic carcinoma and in rhabdosarcoma.630−632 On the contrary, lower glypican-5 expression is related to tumor progression in hepatocellular, prostate, and lung cancer.633−637 Lower expression of glypican5 is a poor prognostic indicator related to lower survival rates in patients with prostate and nonsmall cell lung cancer.634,635 Targeting Interventions for Glypicans in Malignancies. Targeting of glypicans with mAbs, CAR-modified immune cells, and vaccination have been utilized as treatment modalities in malignancies. The administration of monoclonal antibodies against glypican-1 alone638 or conjugated with cytotoxic agents, such as monomethyl auristatin F,606 demonstrate significant antitumor effects in xenograft models of esophageal squamous cell and uterine cervical cancer, respectively. Monoclonal antibodies developed against glypican-2 and conjugated either with pyrrolobenzodiazepine614 or with immunotoxins613 efficiently kill neuroblastoma cells and inhibit tumor growth in vivo. Similarly, the development of CAR T-cells targeting glypican-2 markedly eliminates tumors in a metastatic neuroblastoma mouse model.613 Humanized high-affinity antibodies against glypican-3 have been developed and tested in HCC that inhibits tumor growth in mice.639 Another strategy is the development of high affinity antibodies that recognize a functional epitope of glypican-3 and inhibits Wnt signaling fused with the protein synthesis inhibitory domain of the Pseudomonas exotoxin, which effectively inhibits hepatocellular carcinoma progression.640,641 In addition, bispecific antibodies targeting glypican-3 and CD3 have been used to redirect T-cells against glypican-3 positive tumors.642,643 Glypican-3-specific CAR T-cells as well as glypican-3-specific CAR natural killer cells have been developed to target hepatocellular carcinoma.644,645 GlypicanAG

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

4-amino-quinolyl-6-carbamide (surfen) that antagonize HS bind the cell surface HSPGs inhibiting attachment, absorption, and replication of various viruses.689−692 Several bacteria induce syndecan-1 shedding from cell surface to utilize the ability of syndecan-1 ectodomain to inhibit the antibacterial mechanisms of neutrophils and antimicrobial peptides.681 Protamine, a positively charged protein rich in arginine residues, protects host cells infection by Pseudomonas aeruginosa through inhibition of HSPG shedding.693 Syndecans are implicated in experimental autoimmune encephalomyelitis (EAE), a useful animal model of inflammatory demyelinating disease of the central nervous system. Syndecan-1 is expressed on epithelial cells of the choroid plexus, where CCL20-induced leukocyte recruitment initiates. Shedding of syndecan-1 occurs during the course of EAE and is associated with loss of cell surface-associated CCL20 and increased expression of IL-6. Lack of syndecan-1 in the EAE mouse model results in early leukocyte recruitment, elevated IL-6 expression, and increased disease severity.694 The utilization of surfen ameliorates disease severity by modulating the expression of syndecans.695 The development of novel syndecan inhibitors, like synstatin and ion channels-inhibitors, will extend the biological roles of syndecans in pharmacological targeting that will further improve the efficacy of established therapeutic approaches. 6.4.4. Compounds Interfering with the Functions of Cell-Surface PGs Functions. LMW Hep, modified Hep derivatives, and HS-mimetics display potent anticancer activity via several mechanisms including inhibition of cell surface HSPG-signaling or abrogation of tumor cell adhesion, spreading, and angiogenesis.60,696,697 Another therapeutic intervention for cancer treatment is the utilization of β-Dxylosides as artificial initiators of GAG biosynthesis. Xylosides diffuse within cells and initiate GAG biosynthesis in the Golgi apparatus resulting in the secretion of xyloside-primed GAGs that compete with PGs for binding to various proteins affecting various biological functions. In addition, due to their competing nature with core proteins for biosynthetic enzymes, PGs with reduced or without GAG chains are secreted or bound to the cell surface.129 Treatment with xylosides reduces glycosylation of cell surface PGs and leads to reduced angiogenesis and decreased metastasis by affecting growth factor and chemokine signaling in tumor and endothelial cells.129,134,698−700 Furthermore, it decreases the ability of tumor cells to uptake exosomes that potently stimulate their aggressive phenotype.133

Syndecan-1-mediated activation of IGF-IR, VEGFR, and integrin signaling can be also inhibited by the usage of peptides called synstatins. The coupling of syndecan-1-VEGFR-α4β1 integrin on the surface of myeloma and endothelial cells activates the invasive phenotype and angiogenesis. This complex formation is abrogated by using synstatins based on the binding motif of syndecan-1.679 Syndecan-1 also forms complexes with IGF-IR on myeloma cell surface to prevent their apoptosis. A novel synstatin specific for myeloma cells inhibits the capture of IGF-IR by syndecan-1 and potently inhibits tumor growth and angiogenesis in myeloma tumor xenografts (Figure 11).510 Syndecan Targeting in Inflammatory and Neurodegenerative Diseases. Adequate and timely recruitment of leukocytes to the sites of infection or injury is essential for tissue repair. An imbalance in leukocyte response leads to undesired tissue injury. Syndecans constitute major components of the endothelial glycocalyx, and they have been recently demonstrated as cellular receptors of serine protease inhibitors, such as antithrombin, a single-chain glycoprotein which is a plasma serine protease inhibitor with significant roles in the modulation of blood coagulation. Under physiological conditions, binding of antithrombin to endothelial syndecans (i.e., syndecan-4) downregulates inflammatory response and may induce apoptosis of endothelial cells by disrupting cell− matrix interactions.680 Syndecan-1 is involved in leukocyte recruitment in noninfectious inflammatory diseases via multiple mechanisms.681 Experiments in syndecan-1 knockout mice demonstrated that syndecan-1 inhibits leukocytes adhesion to activated endothelial cells and their migration possibly by interfering between the interaction of leukocyte integrins with endothelial ICAM-1 and VCAM-1.537,681 Furthermore, lack of syndecan-1 evokes expression of pro-inflammatory molecules.537,681,682 Syndecan-1 is important to maintain intestinal mucosal barrier function preventing bacterial translocation.683 Increased syndecan-1 shedding from intestinal epithelial cells represent a novel diagnostic marker for patients with inflammatory bowel disease.684 Syndecan-1 also plays a protective role in experimentally induced colitis, in agreement with reduced syndecan-1 in ulcerative colitis.685 Syndecan-1 knockout mice exhibit impaired wound healing, extended leukocyte recruitment, elevated expression of pro-inflammatory molecules, and increased lethality.685 Syndecan-1 generates chemokine gradients in inflammatory diseases. The interaction of chemokines with HS chains of syndecan-1 gathers chemokines to endothelial cell surfaces at sites of inflammation and creates chemotactic gradients that lead the directional migration of leukocytes.681,686 Similarly, shed syndecan-1 from the surfaces of epithelial cells directs the transepithelial migration of leukocytes.681,687 Shed syndecan-1 is responsible for resolving chemokine gradients by removing sequestered chemokines from inflammatory sites thus preventing prolonged leukocyte inflammation and tissue damage.681,688 In contrast, syndecan-1 is an attachment receptor for a variety of pathogens including hepatitis E virus, papilloma virus, Herpes simplex virus, human immunodeficiency virus, and the use of agents that modulate the interaction of HS chains with virus receptors may represent a novel treatment modality.60,681 Exogenously added Hep, HS chains, and HS mimetics efficiently inhibit infection of host cells by a wide range of HS-binding pathogens.60 Novel antiviral compounds containing the N,N′-bis-5-pyrimidyl moiety as well as a dispirotripiperazine derivative (DSTP 27) and bis-2-methyl-

7. INTRACELLULAR PROTEOGLYCANS: FUNCTIONAL ROLES OF SERGLYCIN Serglycin Regulates Immune System Response and Tumorigenesis

Although intracellular PGs constitute a distinct subfamily, serglycin is the sole member of this class. Serglycin is stored within cytoplasmic secretory granules and vesicles playing important roles in their maturation. Serglycin is essential scaffold for the storage of various components including growth factors, cytokines and proteases, as well as for secretion and target delivery.6,29 Serglycin consists of a small core protein holding a serine-glycine rich repeat that is modified with up to eight Hep/HS, CS, or DS chains (Table 3). Serglycin is widely expressed by hematopoietic cells, although it has been shown that other cell types, such as smooth muscle AH

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

cells, fibroblasts, and endothelial cells, also express serglycin. Over the last decade, it has been shown that serglycin mediates the physiopathology of several malignancies, as it is upregulated in several cancer types including multiple myeloma, acute myeloid leukemia, breast, lung, prostate, colon, hepatocellular, and nasopharyngeal carcinoma, follicular dendritic cell-sarcoma, and gliomas.701−711 Increased expression of serglycin is related to poor prognosis in hepatocellular and nasopharyngeal cancer.708,709 Ectopic expression of serglycin promotes cancer cell growth and invasion in a CS-dependent manner.705 Serglycin secreted from malignant cells such as myeloma and breast cancer cells carries CS-4S chains consisting of approximately 90% GlcAGalNAc(4S) unit. This serglycin variant has increased affinity for collagen type I, the collagenous domains of C1q and mannose binding lectin (MBL) to inhibit the classical and lectin complement pathways, respectively, protecting tumor cells from complement system attack.705,712,713 It is worth noting that only Hep and oversulfated CS-E exhibit similar effects as the CS-4S chains of serglycin for binding C1q and MBL to inhibit complement system during inflammation.712 Notably, CS chains enriched in GlcA-GalNAc(4S) units are attached to PGs, including serglycin in various malignant tissues evoking integrin signaling to promote tumor growth and spread.714 Serglycin induces EMT and chemoresistance as well as enhances the biosynthesis of proteolytic enzymes in breast cancer cells. It also regulates the secretion of IL-8 by breast cancer cells, activating an autocrine IL-8/CXCR2 signaling axis and downstream signaling pathways including PI3K, Src, and Rac to promote breast cancer cell growth and spread.715 Serglycin evokes breast cancer progression by activating CD44/CREB1 signaling to enhance the secretion of TGF-β2 and EMT.706 It activates CD44/nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB)/ claudin-1 and MAPK/β-catenin signaling to promote EMT and chemoresistance in nonsmall lung cancer cells and nasopharyngeal carcinoma.707,716 Serglycin regulates tumor angiogenesis by stimulating pro-angiogenic factors such as VEGF and HGF in in vivo models.701,717 Ablation of serglycin controls the formation of lung metastasis most likely due to the decreased expression of chemokines such as CCL2 in MMTVPyMT-driven mouse breast cancer model.718 Serglycin overexpression also regulates the protein cargo loading of tumorderived exosomes and thus their ability to trigger an invasive phenotype in target cells.719 Although serglycin is constitutively secreted, it is present on the myeloma cell surface promoting their adhesion to collagen type I that enhances MMP-2 and MMP-9 secretion.713 Serglycin plays an important role in the immune system by regulating the cargo of secretory granules of immune system cells.29,720 For example, lack of serglycin severely compromises T- and NK cell cytotoxicity.721 Serglycin is essential for granule maturation as well as perforin and granzyme B storage in cytotoxic T-cells and mast cells thus regulating their activity.721,722 Serglycin-knockout mice are more susceptible to T. spiralis infection by way of decreased mast cell recruitment and diminished protease and inflammatory cytokine release.723 Serglycin expression is increased in LPSinduced inflammation in chondrocytes and regulates the biosynthesis of pro-inflammatory mediators including TNF-α, IL-1β, IL-6, and MMP-9 in a CD44-dependent manner.724 Recent evidence supports the view that the deeper understanding of targeting serglycin, which is associated with

the aggressive mesenchymal phenotype of various cancer types, and serglycin-dependent mediators will help the development of novel therapeutic interventions for cancer therapy.

8. REGULATION OF EXOSOME BIOGENESIS BY SYNDECANS AND HEPARANASE Exosomes have emerged as important regulators of tumor and host cell behavior. Syndecans and their modification by HPSE play a key role in driving exosome biogenesis in tumor cells. Syndecans are endocytosed into endosomes where, via their cytoplasmic domains, they dock with the PDZ domains of syntenin. The formation of a molecular complex with Alix, an event that stimulates exosome budding from the endosomal membrane, will be addressed in this section. The new concept of activated heparanase residing within the endosomal compartment to trim the HS chains of syndecans, a mechanism that facilitates syndecan-syntenin-Alix complex formation to enhance exosomal biogenesis and budding rates, is presented schematically and discussed. 8.1. Syndecans Promote Syndecan/Syntenin/Alix Complex Formation

Extracellular vesicles (EVs) refer to two types of particles released by cells. Large particles of 150 to 1000 nm are referred to as microvesicles (MVs).725 These derive from budding of the plasma membrane and are subsequently released into the extracellular space. In contrast, particles of 30−100 nm are known as exosomes and are, by definition, derived from endosomal compartments. EVs (both MVs and exosomes) can travel locally or distally within the fluid compartments of the body, subsequently dock with recipient cells, and deliver their contents (e.g., proteins, lipids, and nucleic acids). A multitude of wide-ranging functions for these cell-derived vesicles exist, including the regulation of immune cell function, control of cell signaling, transcriptional regulation, and establishment of premetastatic niches.725 Together, these functional studies have cemented the importance of EVs in intercellular communication and identified them as potential therapeutic targets. EV research has further been fueled by the rich source of biomarkers they provide for early prognosis and staging of diseases as a guide for precision medicine.726 All eukaryotic cells, normal or diseased, can release EVs. Evidence indicates that secretion of EVs may be elevated under some disease states, although this remains somewhat controversial.727 There are currently several challenges for researchers studying EV biology. As there are no single molecular markers that discriminate between MVs and exosomes, particle size, buoyant density, and protein composition are commonly used as identification parameters.728 Protein composition is particularly important because it identifies components within the endosomal compartment involved in exosome biogenesis. Another challenge is identifying the cell of origin once an EV is secreted by a cell. This is particularly important when studying the functional impact of EVs in disease. For example, isolation of exosomes present in the serum of a cancer patient will contain exosomes secreted by tumor cells, as well as exosomes secreted by cells within the tumor microenvironment or by other cells throughout the body. Exosome biogenesis occurs via the endocytic pathway (Figure 13). Primary endocytic vesicles fuse to form an early endosome that eventually transitions into a late endosome.729 Next, intraluminal budding of the endosomal membrane AI

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

budding and intraluminal vesicle formation are a topic of intense investigation. Primary regulators of this process are members of the endosomal sorting complex required for transport (ESCRT). The ESCRT machinery facilitates endosomal membrane budding and scission of the newly formed vesicles. To date, approximately 30 proteins have been found to associate with various ESCRT complexes (ESCRT-0, -I, -II, and -III) that act sequentially during vesicle formation.733 While studying MCF-7 breast cancer cells and utilizing a yeast two-hybrid assay, Baietti and colleagues discovered that the Nterminal domain of syntenin interacts with Alix (Bro1/ALG-2interacting protein X), a well-characterized ESCRT complex protein.734 This binding occurs via the LYPX(n)L motif in the N-terminal region of syntenin. A direct interaction between syntenin and Alix was confirmed by coimmunoprecipitation and by Biacore analysis. Syntenin via its PDZ domains also binds with high affinity to a conserved region present within the cytoplasmic domain of all four members of the syndecan family.735 Further analysis utilizing GST-Alix exposed to syndecan:syntenin complexes revealed that syndecan, syntenin, and Alix formed a tripartite complex, thereby linking syndecan to the ESCRT machinery complex (Figure 14).734 Syntenin and Alix were frequently present in exosomes,736 therefore MCF-7 cell exosomes were isolated by density gradient centrifugation and probed by Western blotting for the presence of syndecan, syntenin, and Alix. All three copurified in gradient densities consistent with exosome sedimentation. The presence of exosomes within this fraction was confirmed by electron microscopy. Interestingly, only the syndecan C-terminal fragment (CTF) containing the cytoplasmic and transmembrane domain rather than intact full-length syndecan was present in the exosomes, suggesting that during the biogenesis of the exosomes, the extracellular domain of syndecan was cleaved enzymatically and shed. Full-length syndecan was only present in exosomes secreted by cells in which syndecan had been overexpressed. Surprisingly, flotillin1, another protein commonly found in exosomes was not present in the syndecan-syntenin-Alix exosomes, suggesting that these “syntenin exosomes” represent a distinct class of exosomes. Overexpression of syntenin in these cells resulted in a 2-fold increase in the number of exosomes secreted. Reducing by ∼50% the expression of either syndecan, syntenin, or Alix resulted in an ∼50% reduction in the number of exosomes secreted.734 Together these data demonstrated an important role for the syndecan-syntenin-Alix complex in driving exosome biogenesis. Scrutiny of exosome biogenesis in these cells revealed that the syndecan-syntenin-Alix complex was required for intraluminal budding within the endosome and that the production of syntenin exosomes required multiple members of the ESCRT complex. Subsequently it was discovered that, at least in MCF-7 cells, for syndecan and syntenin to support endosomal budding, they must be phosphorylated by the cytoplasmic tyrosine kinase SRC.737 This phosphorylation occurs on syndecan at the tyrosine residue present in the juxtamembrane DEGSY motif and on syntenin at Tyr 46. Consistent with the requirement for syndecan phosphorylation, it was later found that glypican, which lacks a cytoplasmic domain and thus is not phosphorylated, is not involved in exosome biogenesis.738

Figure 13. Exosome biogenesis. Primary endocytic vesicles arising at the cell surface fuse to form early endosomes. Events in late endosomes lead to formation of intraluminal buds that invaginate into the lumen of the endosomal membrane. In some, but not all vesicles, budding is driven by ESCRT protein complexes. Cytoplasmic contents including proteins, lipids, and nucleic acids enter the buds, and when scission of the membrane occurs at the neck of the bud, they become trapped as cargo within the intraluminal vesicles. Vesicular cargo also includes molecules on the surface of the vesicle, often originating from the cell surface (e.g., cell surface receptors, red). Now containing multiple intraluminal vesicles and referred to as a multivesicular body, the vesicle either becomes degraded within lysosomes or fuses with the cell membrane releasing the vesicles into the extracellular space. These vesicles, now known as exosomes, are available to dock with target cells, deliver their cargo, and influence cell behavior. These exosome-target cell interactions may occur locally, close to the exosome secreting cell, or distally following diffusion of exosomes into the blood or lymphatic system. Additionally, some exosomes become lodged within extracellular spaces where they facilitate cell migration or organization and degradation of the ECM.

initiates the formation of small vesicles that become loaded with proteins, lipids, nucleic acids, and other components of the cellular cytoplasm, some of which are specifically sorted to that location.727 It is important to note that the inward budding of the endosomal membrane results in formation of vesicles having molecular components on their surface that reflect, at least to some extent, those present on the surface of the cell. For example, cell surface receptors will often become localized on the outside surface of the vesicle as it forms within the endosome. Subsequent scission of these endosomal buds at their interface with the endosomal membrane results in the release of multiple small intraluminal vesicles into the endosome compartment. At this point, the endosomes are referred to as multivesicular endosomes or multivesicular bodies. Once formed, the multivesicular endosomes either fuse with lysosomes for degradation or fuse with the plasma membrane resulting in the extracellular release of the intraluminal vesicles. These vesicles are now exosomes.730,731 Interestingly, cells can contain different populations of multivesicular bodies732 raising the possibility that exosomes arising from different endosomes have distinct functions. Given the emerging importance of exosomes in regulating cell behavior, the precise mechanisms governing endosomal AJ

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

extracellular domain from the syndecan CTF also participated in driving exosome biogenesis, perhaps by removing steric hindrance and allowing tighter association of syndecan cytoplasmic domains with syntenin and Alix.734 Shortly after publication of the above work demonstrating a role for syndecans and their HS chains in exosome formation, Thompson et al. reported data revealing that the HS degrading enzyme HPSE significantly enhanced exosome biogenesis.741 HPSE is an endoglucuronidase that cleaves HS chains. It does not completely remove HS from PGs, rather it shortens the chains and in doing so releases fragments of HS 4−7 kDa in size that are known to retain biological activity.742 HPSE is upregulated in cancer as tumors become increasingly aggressive.8 The 65 kDa latent form of the enzyme is secreted by cells and binds to LRP1 or the mannose-6-phosphate receptor for endocytosis and becomes activated intracellularly by cathepsin L cleavage, which generates 50 kDa and 8 kDa fragments that assemble to form the active heterodimeric form of the enzyme.743−745 The impact of HPSE enzyme activity is broad as it can promote angiogenesis, growth, and metastasis of tumor cells.8,746 These activities of HPSE are mostly due to its HS-degrading activity, although HPSE does have some nonenzymatic activities that regulate cell signaling.747 Additionally, HPSE upregulates genes known to promote an aggressive tumor phenotype including VEGF, HGF, and MMP-9.748−751 The finding that HPSE enhanced exosome biogenesis came initially in human myeloma cells transfected with the cDNA for HPSE.741 This resulted in an ∼4-fold increase in HPSE expression by the cells and stimulated an ∼ 6-fold increase in the amount of exosomes secreted. The addition of recombinant, enzymatically active HPSE to myeloma cells enhanced exosome biogenesis in a dose-dependent fashion thereby confirming the findings in transfected cells. Similarly, recombinant HPSE enhanced exosome secretion by MDAMB-231 human breast cancer cells, indicating that the effect of HPSE on stimulation of exosome biogenesis was not cell-type specific. Myeloma cells transfected with a cDNA coding for a mutated, enzymatically inactive form of HPSE failed to enhance exosome secretion. Thus, the stimulatory effect that HPSE has on exosome biogenesis is related to its ability to enzymatically alter the HS chains of syndecan. This was confirmed in a subsequent study utilizing cells in which synthesis of HS was reduced to low levels. Addition of HPSE to those cells had no effect on exosome biogenesis.738 Mechanistically it was determined that HPSE enhances exosome biogenesis by stimulating endocytosis of syndecans and by acting through the syntenin-Alix pathway to stimulate intraluminal budding of vesicles.738 Interestingly, in MCF-7 cells, HPSE did not stimulate the formation of all exosome types, as it had no effect on exosomes containing CD9, CD81, or flotillin-1. In the case of myeloma cells, because syndecan-1 is the only HS proteoglycan detected in appreciable amounts, the effect of HPSE is likely due to its action in shortening the HS chains of syndecan-1. However, in contrast to the effect of HPSE, complete removal of HS from myeloma cells using bacterial HPSE III significantly diminished exosome biogenesis, consistent with the findings Baietti et al. reported in MCF-7 cells.734,741 Analysis of exosome cargo secreted by myeloma cells expressing HPSE at a high level revealed more syndecan-1, VEGF, HGF, and fibronectin than was present in exosomes secreted by cells expressing a low level of HPSE. Syndecan-1

Figure 14. Syndecans, HS, and HPSE drive exosome biogenesis. (1) Syndecan HSPGs localized to the cell surface and anchored by a transmembrane and cytoplasmic domain are endocytosed and reside on the luminal side of early endosomes (2). Molecules bound to the syndecan HS chains on the cell surface such as fibronectin and HS binding ligands (e.g., VEGF, HGF) are translocated into the endosome via their attachment to HS. Enzymatically active HPSE within the endosome modifies syndecan HS chains by cleaving small 5−7 kDa fragments from the chain and leaving chains shortened but still intact. This may remove some of the HS-bound cargo releasing it into the lumen of the endosome. The shortening of HS chains on syndecans by HPSE is thought to enhance formation of multimeric complexes anchored within the cytoplasm through interaction between the syndecan cytoplasmic domain and syntenin. (3) Budding of endosomal vesicles is facilitated by syntenin recruitment of Alix and other ESCRT proteins that together drive the budding and scission process. These processes may be further facilitated by proteolytic cleavage of the syndecan core protein that releases the extracellular domain from the syndecan CTF consisting of the syndecan transmembrane domain and the cytoplasmic domain that binds to syntenin. Shed ectodomains of syndecan still bearing HS chains likely are released as soluble molecules when the multivesicular body fuses with the cell membrane and contents are spilled into the extracellular milieu. In myeloma cells, not all of the syndecan undergoes ectodomain shedding and some syndecan remains on the vesicle surface as an intact proteoglycan. (4) Cargo bound to the syndecan HS when exosomes are released remains functional, and in the case of fibronectin can aid in the docking of exosomes with target cells. HPSE, also bound to HS on the exosome surface, is available for delivery to target cells or to degrade HS within the extracellular matrix.

8.2. Heparanase Trimming of Syndecan HS Chains Enhances Exosome Biogenesis

Oligomerization of syndecans and their clustering within the plasma membrane can be mediated by cargo (e.g., Hepbinding growth factors) that binds to and cross-links the HS chains of syndecans.739 Due to reports that higher-order oligomerization of plasma membrane proteins target them to exosomes,740 the impact of syndecan oligomerization on exosome biogenesis was explored. Blocking enzymes responsible for HS synthesis or treating cells with a bacterial enzyme that strips HS from the syndecan core protein diminished exosome secretion, indicating that HS chains are critical for exosome biogenesis. Cleavage and release of the syndecan AK

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 15. Schematic representation of HA metabolism. HA is synthesized by HAS1, 2, or 3, located in the plasma membrane and use cytosolic UDP-GlcA and UDP-GlcNAc as precursors. HASs catalyze the formation of a HMW-HA that can remain associated with the enzyme (forming the pericellular coat) or released to the ECM. HA can covalently bind with the heavy chain (HC) of interalpha-inhibitor during inflammation. TSG6 catalyzes the transfer reaction of HC to the hydroxyl group of C6 of GlcNAc. Modified HA can form aggregates that appear like cables and possesses the ability to interact and recruit monocytes via CD44. HA fragments can be generated by enzymatic reactions (catalyzed by different HA degrading enzymes) and nonenzymatic reactions (i.e., UV and ROS). LMW-HA can be internalized and degraded in lysosomes. Alternatively, LMWHA can trigger inflammation via activation of TLRs and NF-kB.

targets for blocking exosome biogenesis and function. In fact, recent studies have shown that Roneparstat, a HPSE inhibitor composed of modified Hep, can block exosome docking with recipient cells and the anti-HPSE mAb, H1023, can inhibit exosome-mediated migration of macrophages.752,753 Future studies aimed at a more comprehensive understanding of the role of syndecans and HPSE in regulating exosome biogenesis, composition, and function will likely uncover additional ways in which to target exosomes therapeutically.

was detected on the exosome surface and contained HS chains indicating that in exosomes secreted by myeloma cells at least some intact syndecan is present.741,752 This is in contrast to the report using MCF-7 cells where, as mentioned above, only the CTF was detected in exosomes. The finding that HS chains are retained on the surface of exosomes raises the possibility that cargo bound to syndecans via HS is retained on exosomes and available to mediate intercellular communication when the exosomes dock with recipient cells. In fact, it is likely that VEGF and HGF detected on these exosomes are bound to the exosome surface through interaction with syndecan HS. Interestingly, HS can play a dual role in exosome-cell interaction; HS on the exosome surface captures fibronectin, and HS on recipient cells acts as a receptor for fibronectin thereby docking the exosome to the cell.752 This is consistent with the previously reported role of HS acting as a receptor for exosome docking with recipient cells.133 Moreover, when added to myeloma cells, the exosomes secreted by cells expressing elevated HPSE enhanced cell spreading on fibronectin-coated surfaces and increased the capacity of endothelial cells to invade through Matrigel.741 Taken together, these findings established that syndecans, their HS chains, and HPSE all play important roles in regulating exosome biogenesis, exosome protein composition, exosome function, and exosome docking to recipient cells. It was recently reported that HPSE localizes to the surface of exosomes where it becomes activated and degrades HS within the ECM.753 This provides a mechanism that likely facilitates cell migration and renders HPSE vulnerable for easy transfer to recipient cells. Given the importance of exosomes in intercellular communication in disease states such as cancer and inflammation, syndecans and HPSE represent viable

9. HYALURONAN-METABOLISM, EPIGENETICS, AND FUNCTIONS 9.1. Hyaluronan Metabolism: Overview

HA has the simplest chemical structure of all GAGs, being composed of only GlcA and GlcNAc without further chemical modifications, such as branching, sulfation, acetylation, or epimerization (for structural characteristics see section 3 and Figure 1). Despite this simple structure, HA is ubiquitously produced in all mammalian tissues and is a key component of ECM to influence and regulate many pathophysiological processes. Intriguingly, HA surrounds the oocytes before the fertilization, and HA degradation by hyaluronidases (HYALs) is the first biochemical reaction occurring in mammals’ development allowing sperm/egg interaction and fusion.754 The appearance of HA in nature occurred later than other GAGs (from Tetrapods) supporting the concept that this biopolymer, despite its simple structure with only a space filling function, has an inherent complexity based on its molecular size and interactions with specific receptors and other proteins.755 Thus, HA is critical in normal and pathological biological processes, and its amount is thoroughly AL

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

controlled. Several microorganisms such as Streptococcus uberis, Streptococcus equisimilis, Streptococcus pyogenes, and Pasturella multocida (pathogens of humans and other vertebrates) evolved the ability to produce an HA capsule. This HA has exactly the same structure of vertebrate HA, is nonimmunogenic, and the HA surrounding the bacteria cannot be recognized by antibodies or other specialized immune cells of the host.756 In the human body, HA accounts for more than 15 g in total, located mainly in the skin. The turnover of this polymer is fast and about of 30% of it is replaced every day.757 The fast degradation of HA is carried out by HYALs and hydrolases, which degrade the polymer in small fragments that maintain important biological functions. Usually, HA fragments are internalized by cells and destroyed in the lysosomes. If the cells are unable to properly remove these fragments, their presence can play an important role in several biological processes, including angiogenesis, inflammation, and cell behavior via specific HA receptors.758 Additional factors can affect HA integrity, including free radicals and UV radiation. The fragments of HA can be recognized by CD44 and TLRs triggering specific inflammatory pathways.759−762 Furthermore, HA forms peculiar cable-like structures that are highly adhesive for immune cells via CD44,763,764 giving additional complexity to HA in modulating inflammation765 and other physiological processes, such as ovulation and development (Figure 15).766−768 HA is the only GAG produced outside of the Golgi. HA synthases (HAS 1, 2, and 3) are embedded within the plasma membrane.769 These glycosyltransferases are very peculiar enzymes, since may start the polysaccharide chain without a primer. Recent findings suggest that class I HASs (see below) produce a short chitin-UDP oligonucleotide [(GlcNAc-β1,4)nGlcNAc(α1 → )UDP] as a primer to start the HA synthesis.770 HASs transfer two distinct monosaccharides in different linkages to the growing chain and extrudes it out of the cell. On the other hand, most glycosyltransferases involved in other GAG or glycans synthesis form only one sugar linkage.769,771 In fact, from a structural point of view, HA is a polymer composed of D-GlcA linked to GlcNAc with a glucuronic β(1−3) linkage between GlcA and GlcNAc and a hexosaminic bond β(1−4) between GlcNAc and GlcA (see Figure 1). The disaccharide unit is repeated thousands of times, reaching a molecular weight of millions of Daltons (range from 5 × 105 to 5 × 106 Da).772 The enzymes involved in the HA synthesis have a double catalytic domain as they recognize two different substrates (UDP-GlcA and UDP-GlcNAc) to generate the disaccharide units necessary to create the polymer. The kinetic properties of the HASs are elusive even if extensively studied. The absence of structural information from crystallography maintains many unsolved molecular aspects of these enzymes,769,773 although some similarities can be found with membrane glycosyltransferases, such as cellulose synthase in plants.774,775 HASs are classified in two classes. Class I HASs are multiple transmembrane proteins present in bacteria and animals. Typically, bacterial enzymes contain 6 membrane associated helices, whereas animals HASs have 8 transmembrane domains (Figure 16). Pasturella m. expresses a peculiar HAS belonging to Class II; it has one transmembrane domain at the Cterminal and a completely different mechanism of catalysis.769 Another puzzling question is the presence of three different enzymes in mammals to produce a simple polymer. The

Figure 16. Schematic representation of human HAS2. HAS2, which belongs to Class I, embedded in the plasma membrane while the eight transmembrane helices are represented as cylinders. The large intracellular loop between transmembrane helices 2 and 4 contains the catalytic site. Three critical residues modified by phosphorylation, O-GlcNAcylation, and ubiquitination that modulate protein activity, stability, and dimerization are labeled and shown in different colors.

different enzymes may be necessary to produce chains with varying length, as HAS1 and 2 produce longer polymers than HAS3.776 Further, the catalytic properties among the HASs are different and may justify the three different mammalian enzymes. HAS2, for instance, has several covalent modifications including phosphorylation,777,778 O-GlcNAcylation,779 and ubiquitination (Figure 16).780 HAS3 activity is regulated by trafficking to the cell membrane via Rab10.781 As a key ECM component, HA has important structural properties in tissue biomechanics. Reducing HA content in the skin, a natural event during aging, is responsible for a decrease in skin volume and elasticity, resulting in wrinkles and inelastic, fragile tissue. These findings imply that HA viscoelastic and mechanical properties strongly depend on size and concentration of the polymer. HA polymers have hydrophilic properties that strongly regulate water content in tissues. To better understand the HA capacity to maintain tissue hydration, it may considered that 1 g of HA could bind up to six liters of water.782 However, HA is not just a passive participant in ECM metabolism. HA shows remarkable mechanical properties, ranging from a space filler to a molecular sieve. It is a key molecule with incredible biological properties that posit this GAG among the most biologically active and important molecules in the body. Thus, it is unsurprising that HA performs critical roles in development,783 wound healing,784 cell migration,785 and proliferation,778 nevertheless it is also involved in cancer and metastasis,786 vascular diseases,787,788 and diabetes.789,790 9.1.1. Control of Hyaluronan Synthesis. In mammals, the three enzymes involved in HA synthesis are encoded on different chromosomes and may have arisen from ancestral gene duplication.791 The enzymes can be modulated by covalent post-translational modifications including phosphorylation, ubiquitination, and O-GlcNAcylation. HAS2 activity is modulated by high levels of UDP-GlcNAc, not only because UDP is a precursor but also that HAS2 is O-GlcNAcylated. The O-GlcNAcylation covalently links N-acetylglucosamine (O-GlcNAc) to the hydroxyl group of a serine or threonine side chain. O-GlcNAcylation was found to be linked to cellular features relevant to metastasis, suggesting that higher OGlcNAcylation levels may be a marker for malignancy. These interesting data are presented and critically addressed in this section. A large body of literature suggests that different growth factors and cytokines regulate the HAS expression,792−795 AM

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 17. (A) Schematic representation of the metabolism of dietary monosaccharides (in red) in the formation of activated sugar nucleotides used in biosynthesis of various glycans. UDP-sugars used for GAG and PG biosynthesis are shown in green. All reactions take place in the cytosol with the exception of the reaction in blue outline. In this scheme, almost all the enzymes have been omitted with the exception of the critical enzymes involved in UDP-GlcUA synthesis (UGPP and UGDH), UDP-GlcNAc synthesis through the hexosamine biosynthetic pathway (GFAT), and UDP-GalNAc (epimerase) synthesis. These latter UDP-sugars will be further chemically modified in HS/Hep, CS, and KS biosynthesis. (B) Conversion of UDP-glucose to UDP-glucuronic acid by UGDH. It is highlighted the involvement of 2 NAD in the oxidation of the C6 of glucose with the generation of 2 NADH. The regeneration of NAD via mitochondrial metabolism is critical to supply energy.

downstream of specific cell surface receptors.796,797 It is remarkably important to note that the capacity to regulate enzyme activity is quite fast.798 The diversity among these enzymes is related by their different and unique regulatory mechanisms of synthesis. In most cases, the information is still elusive and is mainly pertinent for HAS2. The HASs are activated in the plasma membrane using cytosolic UDP sugars as precursors (UDP-GlcA and UDP-GlcNAc). The enzyme kinetic is remarkable fast, and the enzyme produces large amounts of polymer in few minutes, supporting the rapid, daily turnover in the tissues. The regulation of HA synthesis is incomplete, complex, and includes several possible mechanisms. The first level of regulation is substrate availability. The substrate for glycosyltransferases used to synthesize all glycans (GAGs, glycoproteins, and glycolipids) are nucleotideactivated sugars. UDP is the most common nucleotide used to activate sugars in animal cells, as it is linked to Glc, Gal, GlcNAc, GalNAc, GlcA, and Xyl. Only two additional nucleotides are used, GDP that is linked to Man and Fuc and CMP that is linked to Sia (Figure 17).799 To note, all the substrates for glycoconjugation reactions are sugar nucleotides generated mainly in the cytosol with the exception for the synthesis of UDP-Xyl and CMP-Sia that takes place in the

Golgi and in the nucleus, respectively (Figure 17). Furthermore, these sugar nucleotide precursors are transported in the ER/Golgi by specialized transporters.800,801 This is crucial as it allows two pools of precursors: one in the cytosol and the other inside the ER/Golgi. Although the real concentration of the sugar nucleotides in these two pools is not easily measurable, the concentration of precursors inside the ER/Golgi would be higher than the cytosol as the transporters have low Km values, ensuring an efficient supply in the ER/Golgi lumen. On the other hand, the cytosolic pool could be directly affected by nutrients, as clearly described for the concentration of UDP-GlcNAc.802 The critical role of the amount of UDP sugars in the HA synthesis regulation is confirmed by experiments using 4-MU, a drug that binds the UDP-GlcA and thereby reduces the bioavailability of this precursor.785,803 Other data stressed the importance of cellular energy homeostasis and HA synthesis. The concentration of the second UDP sugar precursor (UDP-GlcNAc) has an important role but via a different mechanism. The UDP-GlcNAc influences HA synthesis not only by triggering it but also since HAS2 itself is a target for OGlcNAcylation, a protein covalent modification described in 1984 by Torres and Hart.805 Indeed, the O-GlcNAcylation is widely described in several physiologic and pathologic AN

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

conditions, including cancer and chronic diseases.806,807 The cytosolic levels of UDP-GlcNAc are finely regulated by the hexosamine pathway. When UDP-GlcNAc is elevated in the cytoplasm, it induces the enzymatic activity of O-GlcNAc transferase (OGT), which catalyzes the β-O-linkage of one residue of N-GlcNAc to serine 221 of HAS2.779 Similar to AMPK regulation, O-GlcNAcylation is specific for HAS2 and does not affect HAS1 and 3 activities or the synthesis of other GAGs. Protein O-GlcNAcylation is related to the overall metabolic conditions of the cell, and UDP-GlcNAc is defined as “general sensor” of energy level in mammal cells.808 In this case, O-GlcNAcylation of HAS2 substantially increases half-life at the membrane as a mechanism to regulate HA synthesis. HAS2 is usually active on the cell membrane for 17 min, whereas after O-GlcNAcylation, the enzyme can remain active on cell membrane for more than 5 h, increasing the HA content in the ECM.779 Enzyme stability, turnover, and HA production also control HAS2 activity via ubiquitination of Lys190 of HAS2.780 Indeed, proteasome inhibition reduces HAS2 turnover and increases HA production.779 From this point of view, it appears evident that HA content also depends on the enzyme stability on cell membrane.809 It is noteworthy that all covalent modifications are ascribed in HAS2, which is the most common enzyme in most mammalian cells. Interestingly, HAS2 ubiquitination critically regulates HAS2/HAS2 and HAS2/HAS3 complex formation which represents the active form of the enzyme in the membrane.780 Importantly, other enzymes involved in HA synthesis could be affected by different conditions; for instance, HAS1 seems to be important in hyperglycemia and in the early phase of development.810 HAS3 is involved in the formation of microvilli structures in cancer cells.811 Some data shed light to the role of the hyperglycemic conditions and their influence on HAS expression, but the mechanisms are still unclear.812 Other signaling pathways can influence the HA synthesis; for instance, it is known that ERK is able to increase the activity of all three HASs, probably by protein phosphorylation at different residues which are targeted by AMPK.778 In different tissues, the expression of HASs at transcriptional level is finely regulated by cytokines and microenvironment alterations.813 For instance, in inflammation induced by cytokines or ER stress, HA synthesis increases.814 Interestingly, ER stress induced by tunicamycin has the capability to activate HA synthesis in purified ER/Golgi membranes highlighting another regulatory point of HAS via the secretory pathway.785 These data suggest that HA synthesis could occur in ER/Golgi compartments and such “preformed” HA could be efficiently released into the extracellular space upon vesicular fusion with the plasma membrane. Such speculation could be supported by intracellular HA, which may derive from HA catabolism and intracellular synthesis. In a similar context of inflammatory behavior, oxidized low density lipoprotein (LDL) and 22-oxysterol increased HAS2 and HAS3 levels in an in vitro atherosclerotic model.815 Modifying the lipid microenvironment around HASs, and probably the transmembrane helices, can interfere with the enzymatic activity. It has been revealed that cholesterol and cardiolipin can further modulate HA synthesis.816,817 Another critical aspect in HA synthesis regulation that remains largely unknown is how HASs control the length of the polysaccharide chain they synthesize. The UDP-sugar precursors may have a pivotal role in determining the

molecular mass of the growing HA polymer as well as the presence of small HA oligomers that can work as acceptor sizes in purified bacterial Pm HAS enzymes.818,819 Recently, experiments were conducted on SeHAS. This protein, as in mammalian enzymes, possess Bx7B repeats at the C-terminus and has a critical role in determining HA size. Mutations in such motifs alter the chain length without altering the synthesis rate.820 Although there are no experimental evidence, it is possible that post-translational modification of such motifs could modulate HA size in vivo under different conditions. 9.1.2. Energy Supply and Hyaluronan Synthesis. The regulation of the HA synthesis is critical and includes different mechanisms which are still poorly understood. One of them is the substrate availability, and the number of UDP-sugars substrates can influence HA synthesis. Altering UDP-sugars availability in the cytoplasm increased HA production in primary cells with increased expression of HAS2 and HAS3. HA synthesis depends also on the AMPK activity, as discussed above. This regulation only affects HA production and not the synthesis of other GAGs. The data correlate energy level with the ability of cells to produce HA and are presented below. The activity of adenosine monophosphate activated protein kinase (AMPK), the master sensor kinase of cellular energy level, regulates HAS2 activity in mammalian cells.804 The activation of AMPK by AICAR induces the phosphorylation of HAS2 at Thr 110, blocking the enzyme activity.804 Confirming the specific regulation of HASs, AMPK activity acts only on HAS2 and not on HAS1 and 3. Moreover, the activity of AMPK plays a role only for the HA synthesis and not for the synthesis of other GAGs, indicating that the HA is the only GAG affected by the energy modulation, thereby aligning cellular bioenergetics with HA synthesis. UDP sugar precursors are molecules with a high-energy cost for the cells, competing with glycolysis and other catabolic pathways for their synthesis. Hence, GAG synthesis is possible in tissues with a good oxygen supply, as two oxidative reactions are necessary for UDP-GlcA synthesis (Figure 17). The synthesis of UDP-GlcA is a critical step for all GAGs, except KS, which does not contain hexuronic acid. In fact, KS is usually common in tissues with poor oxygen supply or even without a vascular system. The synthesis of UDP-GlcA requires the action of the UDP-glucose dehydrogenase (UGDH), which produces UDP-GlcA from precursor UDP-Glc. This reaction is possible in the presence of NAD, which is transformed to NADH during the double oxidation of the C6 of UDP-Glc. This uncommon reaction has a remarkable role in terms of cellular energy balance. From this point of view, the costs of UDP-GlcA synthesis are completely balanced by the reoxidation in mitochondria of the two NADH molecules produced by the UDP-GlcA synthesis. The stoichiometry of the reaction of HA synthesis indicates that the disaccharide units contain one molecule of GlcA and one of GlcNAc with a ratio of 1:1. The five ATPs obtained by the reoxidation of two molecules of NADH in mitochondria repay the energy cost of the synthesis of the nonsulfated backbone of HA. UDP-sugars availability has a critical role on the HA synthesis; in fact, the modulation of UDP-GlcA availability by overexpressing or silencing UGDH in cytoplasm can dramatically influence the HA production as well as the expression of HAS2 and 3.809,821 This aspect is confirmed by using the 4-MU, a drug that binds UDP-GlcA and thereby reduces the bioavailability of this precursor. AO

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

that HYALs activity might be part of a finely tuned system which includes HA synthesis and degradation.836 Species of ROS can cut the HA chain generating biological active fragments. Usually ROS are present in the injured tissue, in inflamed areas and in tumor microenvironment. They may provide a mechanism for generating HA fragments in vivo and may further exaggerate the inflammatory state as HA fragments have shown the significance of HA size during disease. Another intriguing aspect is related to the internalization of fragments inside the cells and their degradation into monosaccharides in lysosomes. The identification of a HYAL responsible for the initial degradation of HA on the cell surface is still elusive, and the interaction of these enzymes and HA receptors, such as CD44, is another critical aspect. Recent identification of TMEM2 as a cell surface protein acting as a potent HYAL suggests that it may be possible that the “missing” cell surface HYAL is not identified and that novel models of HA catabolism should include the presence of this protein.759 In theory, HA oligosaccharides could be classified as matrikines as they have all characteristics of bioactive ECM fragments.837 Oligosaccharides produced by the catabolic activity of HYAL, if not rapidly internalized by the cells, can diffuse through tissues and bind HA receptors on adjacent cells, acting as intracellular signals such as NF-κB and ERK. 9.1.4. Epigenetics of Hyaluronan Metabolism. Aside from the post-translational modifications that act to modulate HAS2 activity, enzyme turnover, and gene regulation via transcription factor recruitment to the HAS2 promoter; recently it has been discovered that HAS2 is regulated by epigenetic mechanisms, such as microRNAs (miRNAs) and chromatin modifications. These findings are very intriguing as epigenetics can be influenced by environmental factors, such as nutrition, drugs, and pollutants. One of the main epigenetic regulators of HAS2 expression is the long noncoding RNA (lncRNA), HAS2-AS1.838,839 This RNA belongs to the natural antisense transcript (NAT) as it is transcribed on the opposite DNA strand with respect to HAS2 and shares the first exon of HAS2 (Figure 18).840 Such lncRNAs are transcribed from specific loci in the genome but are not translated into proteins. Erroneously, lncRNAs have been classified as byproducts of transcription without any identifiable role; however, their importance and specificity have

9.1.3. Hyaluronan Degradation. The amount of HA in the body is strictly regulated by the cells in tissues and via the lymphatic system to be removed in the liver. The size of the polymer is critical for its biological function, and the size in healthy tissue is high molecular weight (HMW) (about 103 kDa). Moreover, very HMW-HA is found in naked mole rat,822 a rodent which shows incredible longevity and cancer resistance. These data shed light on the importance of HA molecular size in senescence and cancer development. The degradation of HA is therefore a key step of HA biology and is obtained by hyaluronidase and external factors. It is hypothesized that HMW-HA in the extracellular space is digested to small fragments by HYALs. They are classified as endo-β-N-acetylglucosaminidases according to their hydrolytic mechanisms.823,824 Six HYALs have been described in humans: HYALs 1−4, HYALP, and PH20, which are all β(1−4) endoglucosaminidases. Human HYALs are encoded by three genes, HYAL1, HYAL2, and HYAL3, coding for HYAL-1, HYAL-2, and HYAL-3, respectively.823 These genes are tightly clustered on chromosome 3p21.3. HYAL-1 and -2 are the major HYALs in tissues. HYAL-2 is a GPI anchored protein with extracellular activity. HYAL-2 acts on HMW-HA generates fragments with a size of about 20 kDa. HYAL-1 appears to be a lysosomal protein to cleave HA into small disaccharides. The role and activities of HYAL-3 is still elusive, and few reports are available on this enzyme as it is reported in experiments based on KO mice.825 Three more genes, HYAL-4, PHYAL1, and SPAM1 (Sperm Adhesion Molecule1), are clustered on chromosome 7q31.3. HYAL-4 is a transcribed pseudogene in humans, and PH-20 digests oocyte HA, facilitating spermatozoa penetration through the cumulus.826 HYALs 1−4 in liver and serum are enzymatically active in acidic environments (pH 3 and 4), whereas PH-20 and other HYALs from insects and snake venom are active at neutral pH.827 HYAL-1 is active intracellularly and is common in mammal tissues. Related to HYAL-1 deficiency is the mucopolysaccaridosis type IX, also called hyaluronidase deficiency. HYAL-2 is a GPI-anchored receptor that operates in an acidic microenvironment at the cell surface.828,829 This enzyme is active in acidic microenvironments degrading HMW-HA into LMW-HA (about 20 kDa), which is internalized and further digested to smaller oligo HA (oligo-HA) by HYAL1.757 The recent therapeutic use of HYALs shows promising applications in the treatment of solid tumors, where the HYAL treatment improves chemotherapy efficacy.830 More recently, Yamaguchi and colleagues described TMEM2, a transmembrane protein with a strong HYAL activity.831,832 TMEM2 digests HMW-HA into ∼5 kDa fragments but cannot cleave other sulfated GAGs, indicating its specificity to HA.832 CEMIP/KIAA1199 is a recently identified HA-binding molecule that has HA-degrading activity (also known as HYBID),833 with the participation of the clathrin-coated pit pathway. Interestingly, this enzyme plays a key role in cancer development as well as in the skin biology and senescence.834,835 Discovery of these new enzymes suggest that HA chemistry is still not completely understood and underscores the critical role of HA in ECM biology. The role of HYAL in human pathology and cancer is still unclear, and several controversial data have been reported. However, these contradictory data indicate that HYAL-1 and HYAL-2 might promote or suppress tumor development, suggesting

Figure 18. Schematic representation of the genomic organization of the HAS2 locus on human chromosome 8. HAS2-AS1 is transcribed on the opposite strand in respect of HAS2 and this lncRNA can have different role to control HA expression. In the nucleus, HAS2-AS1 can recruit enzymes to modify the chromatin or alter chromosome topology whereas in the cytosol HAS2-AS1 can stabilize HAS2 mRNA or act as ceRNAs to compete with miRNA binding. AP

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

been universally accepted. Indeed, lncRNAs have a plethora of molecular functions aiming for gene expression regulation.841 HAS2-AS1 can be recognized in every tissue that coexpresses HAS2 and recent results highlight that this lncRNA is necessary for proper HAS2 expression. There are several ways in which HAS2-AS1 is involved in HAS2 expression, depending on the type of tissue or cell. In vascular smooth muscle cells, HAS2-AS1 is involved in chromatin decondensation around the HAS2 promoter to favor transcription factor binding. Such reorganization of the chromatin could be ascribed to the O-GlcNaclyation of histones with HAS2-AS1 mediating the recruitment of epigenetic modifiers in proximity to the HAS2 promoter.842 As HAS2-AS1 expression is regulated by nutrients such as GlcNAc or glucose, this lncRNA could be critical in diabetes where HA alterations occur in blood vessels.843 The HAS2-AS1 transcript also binds HAS2 mRNA via the complementarity region and stabilizes HAS2 mRNA, favoring translation in renal cells.844 An alternate function of HAS2-AS1 could affect genome organization. Intriguingly, the expression of HAS2 and, its receptor CD44, is always coordinated regulated but without a clear explanation mechanism. Although an explanation could reside in the promoters, it is possible that HAS2-AS1 could alter chromosomal topology that would juxtapose the regulation of both genes, in a similar manner to other lncRNAs, such as HOTTIP845 or HOTAIR.846 Indeed, active genes on the same or different chromosomes are organized into 3D domains that require lncRNAs for proper nuclear architecture.847 In addition, once lncRNAs are translocated in the cytosol, they may exhibit different abilities to regulate expression. One of the most intriguing properties of lncRNA (as well as transcribed pseudogenes) is the ability to bind miRNAs, like a sponge, and compete with the target mRNA for shared binding. Therefore, lncRNAs can work as competing endogenous RNAs (ceRNAs), and this mechanism allows the fine-tuning of gene expression of multiple targets.848 This mechanism has been found also in coding mRNAs and, for example, CD44 regulates CDC42 (a cell cycle regulator) by competing for miR-216, miR-330, and miR-608,830 while Vcan V3, by antagonizing miR-199a, modulates fibronectin.849

Figure 19. HA interacting proteins. Schematic representation of proteins that can bind in a noncovalent fashion to HA. HA is recognized by two classes of proteins containing the Bx7B motif or the LINK module. TLR2 and TLR4 are necessary to induce the proinflammatory response of LMW-HA, but it is still debated whether HA can directly bind to such receptors. HA receptors are shown in green.

contain the LINK module or the Bx7B motif but are necessary to trigger the response after LMW-HA stimulation.761 Although the physical interaction between TLR and HA has not yet been demonstrated, it is possible that the polyanionic nature of HA can mimic the canonical ligands of TLR2/4, such as lipopolysaccharides. HAPLN 1−4 family has been described to interact with HA and PGs, thereby stabilizing such multicomponent complexes.180 The functions of HA in tissues are due to the specific properties of the polymer, which acts as a space-filling and shock-absorbing polymer. HA also interacts with other ECM molecules, such as aggrecan, neurocan, and versican, and these networks act as architectural scaffolds for cells that contribute to the biomechanical properties of the tissues. It is noteworthy to observe that HA activities strongly depend on polymer size. HMW-HA shows antiangiogenic, immune suppressive, and anti-inflammatory activities, and induces tissue reparative process as described in wound healing.761 In contrast, the fragments, called HA oligosaccharides when 20 million Da) that also protects those against cancer.822 On the other hand, HA in the skin is not

9.3. Hyaluronan and Its Biotechnological Applications

Due to its biophysical properties, HMW-HA shows several biological properties including molecular sieving, as a lubricant and space filling activities in connective tissues.882 In recent years, the use of HA in several areas of biomedicine increased dramatically. HA has roles in several therapies in regenerative medicine and in skin, often with cosmetic purposes.883 In wound healing or in surgery as an antiadhesion agent,884 HA shows an effective activity due to its capacity to improve cell growth and angiogenesis. The biotechnological properties of HA depend on its structure. HA chains have β linkages which expose bulky groups (the hydroxyls, the carboxylate moiety, and the anomeric carbon on the adjacent sugar) and for this reason are in sterically favorable equatorial positions, favoring the chemical accessibility. All axial positions are occupied by hydrogen atoms.885 In aqueous solution, HA develops several hydrogen bonds which stiffen the chains. These chemical conditions in solution induce a twisting ribbon structure AR

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Figure 20. An interaction network of Hep/HS. This subnetwork comprises 262 protein-Hep/HS interactions, which were retrieved from the MatrixDB database (http://matrixdb.univ-lyon1.fr/).914 Proteins are identified by their UniProtKB accession number (http://www.uniprot.org/ )915 and multimeric proteins (e.g., collagens and integrins) by their Complex Portal accession number (https://www.ebi.ac.uk/complexportal/ home).916 Proteins and their fragments (ES, endostatin; ER, endorepellin; HepV, fragment of the α1(v) collagen chain; PEX-MMP2, hemopexin domain of MMP-2; NC1-XVIII, C-terminal domain of collagen type XVIII) are color-coded according to their location/function. Blue, ECM (51); green, ECM degradation (10 proteins); pink, cytokines and growth factors (22 proteins); orange, complement system (9 proteins); red, integrins (5 proteins); and black, membrane and intracellular proteins (165 proteins). The network has been visualized with Cytoscape (http://www. cytoscape.org/).917

forming an extended random coil structure. These chains interact with other chains, forming random coil structures called entanglements. In solution, even 1% HA chains can entangle each other forming a molecular network. The

spontaneous process of entanglement starts at 1 mg/mL of HA.886 The helical chain of HA in solution can bind 1000 times its weight in water.887 At 1%, HA forms a jelly acting as “quasi-plastic material” with viscous elastic properties.888 The AS

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

lubricating properties of the polymer and all its physical properties depend on these molecular interactions in water. HA is therefore a perfect lubricant material completely compatible with mammalian cells, and it is largely used in therapy to replace synovial fluid in joints and to prevent in abdominal fibrotic adhesions following surgery. Several hydrogels based on the chemical properties of HA are commercially available as they are completely biocompatible.875 The viscous-elastic properties of HA hydrogels depend on chemistry and concentration of the polymer, increasing the stiffness with the increase in HA concentration and chain size. The rheological properties of HA present in mammals control water content and therefore the ion concentration.889 The HA gels are biodegradable, biocompatible, bioresorbable, and play a critical role in regenerative medicine. HA scaffolds induce cell differentiation and growth890 and are useful properties in wound healing.891−893 HA hydrogel can also be used as an efficient and efficacious drug delivery system.894,895 The polymer can easily be modified with adipic hydrazide, tyramide, benzyl ester, glycidyl methacrylate, thiopropionyl hydrazide, or bromoacetate, either at the carboxylic group of GlcA or at the C-6 hydroxyl group of the GlcNAc.896 In cosmetics, the HA polymer is common in commercially available products due to its moisturizing property. Indeed, the elastic properties of skin improve by the regular application of HA, despite not understanding the biological effects on keratinocytes at the molecular level.897 Nevertheless, even robust scientific data are still necessary to completely understand the HA role in topic cutaneous application. Several sunscreen products containing HA showed important antifree radicals action with protection against ultraviolet irradiation.898 In plastic surgery, HA gels with entanglements or chemical linkages between chains are widely used as filler to treat facial lines and wrinkles.899,900 The success of this application is due to the greater tolerability of HA filler compared to collagen products.901−903 Finally, HA is also present in the nutraceutical market, as beverages, food, and confectioneries have been approved as health food worldwide. Recent advances in nanomedicine have offered new valuable tools for cancer detection, prevention, and treatment exploiting the unique properties of HA as a novel powerful molecule for drug delivery and molecular targeting in cancer.

kinetics, affinity, and expression data in GAG/PG interactomes are also addressed. 10.1. Identification and Characterization of Protein-GAG Interactions

GAGs interact with various protein families,67,69,904 including ECM proteins, membrane proteins, such as integrin receptors,905−907 lipoproteins, enzymes,21 cytokines, including those of the TGF-β family,908 and chemokines, such as neutrophil-activating chemokines (Figure 20).75 GAGs also interact with pathogenic proteins, thereby contributing to host−pathogen interactions and viral entry into host cells.909−912 The Zika virus, which has been declared in February 2016 a Public Health Emergency of International Concern by the World Health Organization, interacts with GAGs as other pathogenic flaviviruses do. Envelope protein E of Zika virus binds to Hep and to brain and placental CS and HS.913 The biological activities of GAGs are mediated by their interactions with proteins. Their binding to ECM proteins contributes to ECM assembly and organization. GAG−protein interactions sequester growth factors in the ECM and contribute to the formation of chemokine and growth factor gradients. GAGs also act as coreceptors, facilitating ligand− receptor interactions, and are involved in cell-matrix interactions, cell adhesion, cell signaling, and cell trafficking. GAG-chemokine interactions mediate neutrophil trafficking (see Figure 11B).75 The most extensively studied protein−GAG interactions are those established by HS and Hep that is frequently used as a HS substitute for in vitro studies.69,918 At least 435 Hepbinding proteins, including ECM proteins, membrane proteins, lipoproteins, enzymes, enzyme inhibitors, cytokines, chemokines, and proteins involved in neurodegenerative diseases, have been identified as Hep/HS-binding proteins.69,908,919 Furthermore, HA interacts with numerous hyaladherins, which comprise cell receptors [CD44, RHAMM, LYVE-1, and the HA receptor for endocytosis (HARE/stabilin-2)], ECM proteins and plasma proteins.758 HA binds to the HAPLNs to form large PG assemblies of hyalectans (e.g., Acan) in the ECM and interacts with CSPGs in cartilage and perineuronal nets. Hyaladherins have been studied for targeted drug delivery as HA-drug bioconjugates.920 KS interacts with about 200 proteins, including secreted, membrane, and intracellular partners.921 Indeed, several GAGs may also have intracellular partners. HA has been found to be associated with microtubules, and the mitotic spindle in arterial smooth muscle cells.922 HS has been detected in the nucleus, where it plays regulatory roles related to cell proliferation, transcription, and transport of cargo to the nucleus.923 CS and DS have been detected in the nuclei of HeLa cells and in rat ovarian granulosa cells, respectively.924,925 Deciphering the molecular mechanisms of GAG−protein interactions and their specificity is a prerequisite to better understand the biological processes they mediate and to design inhibitors modulating protein−GAG interactions. Furthermore, the specificity of these interactions determines the diffusion rate and spatial distribution of proteins, as shown for several FGFs in the pericellular matrix of fibroblasts, which is important for cell-matrix interplay.926 10.1.1. Biophysical Techniques. Biophysical techniques, incuding SPR,532,910,927−930 biolayer interferometry (BLI),930 isothermal titration calorimetry (ITC),927,931−933 and microscale thermophoresis (MST),934 are frequently used to

10. PG/GAG INTERACTOMES: TOPOLOGICAL, STRUCTURAL, AND FUNCTIONAL ANALYSIS The interaction networks of GAGs and PGs give insights into the molecular mechanisms they use to regulate many biological processes in a concerted fashion. The integration in these networks of data on GAGs and PGs and their partners (e.g., 3D structures, transcriptomic and proteomic data, and functional annotations) and on their interactions (e.g., binding sites, kinetics, and affinity) results in tissue- and biological process-specific networks and reflects their permanent molecular rewiring, which occurs in vivo, depending on the biological context. The identification of protein−GAG interactions, the biophysical techniques, and computational studies used to characterize protein−GAG interactions and kinetic parameters as well as the GAG microarrays prepared with natural or synthetic GAG oligosaccharides used for highthroughput screening of GAG−protein interactions are illustrated in the sections below. Moreover, the interaction networks of PGs and GAGs and the integration of structural, AT

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

binding proteins. Three sequences enriched in basic AA residues, XBBXBX, XBBBXXBX, and XBBBXXBBBXXBBX, where B is a basic residue and X is a hydropathic residue, have been found in Hep-binding proteins.946 At least 50 AA sequences enriched in basic residues and binding to HS/Hep have been listed.904 The RKQRRERTTFTRAQL sequence of the homeobox protein OTX2 was found to bind to CS C, D, E, and to Hep with the RK doublet playing a major role in the recognition process.931 Polar AA residues, Asn and Gln, are also found in Hep-binding sites, where they contribute to the specificity of GAG−protein interactions.947 A Gly box (XBXGXXBBG), identified in FGFs, has been reported to be also involved in the specificity of Hep/HS-binding.904,948 A motif containing two cationic residues (Arg or Lys) and one polar residue (preferentially Asn, Gln, Thr, Tyr, or Ser), the socalled “CPC clip motif”, is conserved in Hep-binding proteins of the Protein Data Bank.949 A discontinuous Hep/HS-binding site based on basic residues has been located at the tips of the first β-strand finger loops of TGF-β cytokines.908 GAG binding sites may be comprised of several basic clusters fulfilling different functions. IFN-γ for example, contains two clusters, one contributing to the bioactivity of the cytokine and the other being exclusively involved in HS molecular recognition.927 Three AA sequences have been described as HA-binding motifs.950 They consist of the link module (∼100 AA residues), a RR sequence,951 and a sequence of seven AA residues flanked by either an Arg or a Lys residue, the B-(X7)B motif where X is any AA residue except an acidic one.952 However, there is no direct evidence that this sequence mediates HA binding in all HA-binding proteins.176 GAG features contributing to protein binding may be studied with selectively desulfated GAGs, and with GAG oligosaccharides of defined sequences and length, which matter for interactions with proteins. A number of oligosaccharides are now commercially available from Iduron (UK), which eases the investigation of the molecular features of GAG involved in protein binding. GAG conformation and the distribution of sulfate and carboxylate groups contribute to the specificity of GAG−protein interactions.68 2-O, 6-O and NSulfate groups of Hep/HS are involved to a various extent in the interactions with proteins regulating angiogenesis.953 Some proteins bind to common GAG repeat units as shown for the Zika virus envelope protein, which interacts with the major trisulfated disaccharide [IdoA(2S)-GlcNSO3(6S)] found in HS.913 In contrast, Hep cofactor II is activated by a HS hexasaccharide containing two rare 2-O-sulfate GlcA residues.954 Moreover, the 6-O-sulfate groups of HS are involved in the binding of several proteins (growth factors, matricryptins, and adhesion proteins).955 HS 6-O-Sulfation is a dynamic process regulated in humans by two extracellular 6-Oendosulfatases (Sulfs), which postsynthetically catalyze the removal of HS 6-O-sulfate groups.955,956 6-O-Desulfation can regulate GAG−protein interactions in a protein-specific manner, releasing growth factors sequestered in the ECM and modulating cell signaling. It prevents, for example, the formation of the FGF/HS/FGFR ternary signaling complex, whereas it decreases the affinity of the Wnt/HS interaction, leading to Wnt-induced signaling mediated by the Frizzled receptor.955 Cations bind to Hep/HS, influence their conformations, and may modulate protein-Hep/HS interactions.918 Annexin V, a phospholipid-binding protein interacts with HS in a Ca2+-dependent manner,918 whereas Zn2+

characterize protein−GAG interactions and to calculate kinetic parameters (SPR, BLI) and/or the equilibrium dissociation constants (SPR, BLI, ITC, and MST). A combination of SPR imaging (SPRi), SPR and BLI has been developed to build protein−GAG interaction networks.930 SPR is the most widely used to determine the association and dissociation rates and the affinity of GAG−protein interactions as shown in a data set comprising 125 GAG−protein interactions characterized by SPR.910 ITC provides thermodynamic parameters (enthalpy, entropy, and free energy) and stoichiometry of protein−GAG interactions.933 Quartz crystal microbalance with dissipation (QCM-D), another label-free technique, has been used to characterize HS- and HA-protein interactions.739,935 It gives insights into the viscoelastic properties of the immobilized GAGs and their potential rearrangement upon protein binding. Using this technique and HS chains grafted to streptavidinfunctionalized oligoethylene glycol monolayers or supported lipid bilayers, it has been shown that cytokines and growth factors cross-link HS depending on the architecture of their HS-binding sites.739 GAGs may stabilize cytokine oligomerization, which contributes to the affinity and specificity of GAGbinding.929,936 The 3D structure of GAG−protein complexes provides insights into the mechanisms of GAG−protein interactions.937,938 The binding sites of GAGs on GAG-binding proteins have been characterized by performing binding assays with mutants, protein domains, or peptides using the above techniques. A strategy for defining critical AA residues involved in Hep-protein interactions is based on the cross-linking of one Hep-binding protein to Hep beads followed by the proteolytic degradation of Hep-protein complexes and the identification of the peptides by N-terminal sequencing.939 Selective labeling of Hep-binding Lys residues (i.e., acetylation of nonbinding Lys residues while the protein is bound to immobilized Hep and biotinylation of the residual free Lys residues after Hep removal) has been used to identify Hep-binding Lys residues of vaspin (serpin A12).934 Another “protect and label” strategy is based on the protection of Hep/HS-binding residues toward chemical modification by N-hydroxysuccinimide acetate. After dissociation from Hep, the lysine residues involved in Hep/ HS-binding are labeled with N-hydroxysuccinimide biotin. Biotinylated peptides are purified and identified by mass spectrometry after proteolytic degradation.940 Lysine residues involved in GAG interactions can be characterized by a methylation-NMR approach based on reductive 13C methylation of the Lys side chains and 1H−13C correlation experiments.941 NMR has been combined with other techniques to characterize the binding of a cytokine, IFN-γ, and FGF to HS927,942 and of IL-8 to CS.943 NMR may also be used to obtain structural information on GAG partners using 13 C-labeled GAG oligosaccharide of defined sequences as shown for the interaction of a chemokine (CXCL12α) with a HS-like oligosaccharide.944 Amide hydrogen/deuterium (H/ D) exchange mass spectrometry, in combination with molecular modeling and docking experiments, allows the study of binding interface of protein−GAG complexes.945 Computational approaches and GAG-arrays are also helpful to decipher the molecular mechanisms of GAG−protein interactions as described below. The mechanisms of Hep/HS-protein interactions have been recently reviewed.69,904 Conformational changes in protein upon GAG-binding may be assessed by fluorescence spectroscopy.928,943 Specific motifs have been identified in GAGAU

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

linker and has been implemented in the first dedicated automated oligosaccharide synthesizer (Glyconeer 2.1). This approach has been adapted to the synthesis of GAG oligosaccharides, which requires sulfation and acetylation. Two hexasaccharides of CS 4- and 6-sulfate have been synthesized as proofs of principle.981 Synthetic CS tetrasaccharides have been printed on aldehyde-coated glass surfaces, whereas CS, DS, HA, Hep, HS, and KS have been printed onto poly-DL-lysine-coated glass surfaces.974 HS hexasaccharides and heptasaccharides bearing amino groups at their reducing ends have been printed onto an N-hydroxysuccinimide ester functionalized glass slides and covalently immobilized via amide bonds.975,979 Hep octasaccharides have been also spotted onto γ-aminopropylsilane microarray slides.978 HS derivatives immobilized onto aminosilane slides via their reducing ends or onto a hydrazide-derivatized self-assembled monolayer on a gold-coated slide surface can be probed by matrix-assisted laser desorption/ionization time-of-flight mass spectroscopy, SPR, and QCM.982 GAG oligosaccharides can be modified before spotting on the microarray surface. They have been tagged with lipids and immobilized noncovalently as neoglycolipids on nitrocellulosecoated glass slides.983 The first neoglycolipid-based oligosaccharide microarrays included oligosaccharides from CS 4- and 6- sulfate (dp2 to dp20, dp: degree of polymerization) and DS (dp2 to dp20) and were probed with anti-GAG antibodies and GAG-binding proteins (IFN-γ and the chemokine RANTES).984 GAGs and GAG oligosaccharides have been modified by a pyrrole moiety at their reducing end and covalently immobilized on a gold surface by electrocopolymerization of the pyrrolated oligosaccharides.973 Another method relies on the physical adsorption of full-length GAGs (CS, DS, Hep, HS), LMW Hep and HA, and selectively desulfated Hep on bare gold chips without any chemical modification, during or after the spotting process.906,930,977,980 GAG arrays have been used as screening tools to identify new GAG-binding proteins, such as neurotrophins, which bind to CS-E,974 a fragment of collagen type XVIII, endostatin, which interacts with CS and DS and α5β1 integrin, which binds to HA, Hep, and HS.906 GAG arrays can also be probed with cells to investigate functional GAG−cell interactions. FGF signaling has been measured in HS-deficient cells overlaid on a Hep oligosaccharide array and was shown to depend on the oligosaccharide size. This opens new perspectives for functional glycomics screening with intact living cells.972,976 Protein arrays have been also probed with fluorescently labeled GAGs to find new GAG-binding proteins. About 450 GAG-binding proteins (185 Hep-, 62 HS-, 98 CSB-, and 101 CSC-binding proteins) have been identified using Escherichia coli proteome chips.985

increases the binding of endostatin to HS and its antiproliferative effect on endothelial cells stimulated by FGF-2.957 10.1.2. Computational Studies of Protein−GAG Interactions. Computational approaches including docking, molecular dynamics, and molecular modeling have been applied in combination with experimental methods to identify GAG-binding sites and to build 3D models of protein−GAG complexes. Docking experiments have been used to predict Hep binding sites on endostatin,957,958 FGF-1,959 FGF-2,960 αvβ3 integrin,961 HS-binding sites on IL-8 and growth factors,943,962,963 and CS-binding sites on cathepsin S.928 The interfaces of protein−GAG complexes are more hydrated than those of protein−protein complexes, and water-mediated interactions can be taken into account in docking experiments in addition to electrostatic-mediated interactions.962 Another feature to consider for docking experiments is that iduronic acid, found in Hep, HS, and DS, is either sulfated (2S) or not, and that IdoA(2S) can adopt 1C4 and 2S0 ring conformations, which affects docking solutions and free-energy calculations. In many studies, only one conformation is taken into account, which may result in misleading conclusions. The importance of exploring the conformational space of IdoA in computational studies of Hep/HS-protein interactions has been recently highlighted.964 A molecular dynamics docking method, integrating receptor flexibility and explicit solvation (dynamic molecular docking), has been developed for protein−GAG interactions and applied to the characterization of Hep/FGF-1 interactions.959,965 Protein electrostatic potential calculations are used to define protein surface areas as putative GAG binding sites.958 Another computational method (GAG-Dock) has been set up to accurately predict the binding poses of protein-bound GAGs and applied to the docking of Hep and CS derivatives onto the RPTP-σ and Nogo receptors 1−3.966 Furthermore, a method based on computational modeling (known as combinatorial library virtual screening) allows the design of high affinity and high specificity GAG-binding sequences from GAG combinatorial libraries and GAG-binding proteins. It has been validated using the well-characterized antithrombin-thrombin-Hep complex.967 Computational approaches are also used to build 3D models of GAGs and GAG−protein complexes, to investigate their molecular structure and dynamics, and to characterize protein−GAG interactions at the molecular level.25 Coarse-grained models of HA, Hep, and CS are available,968,969 including heterogeneous CS 4-sulfate and DS 4-sulfate with 200 monosaccharides.25 In addition, molecular modeling of PGs and their GAG chains have been performed at the atomic scale for bikunin and a heterogeneous CS chain and for decorin with a DS or a CS chain.25,970 10.1.3. GAG Microarrays. GAG microarrays displaying natural or synthetic GAG oligosaccharides are used for highthroughput screening of GAG−protein interactions, the determination of the length and chemical groups of GAGs, which are crucial for the interactions, and the study of GAGcell interactions.971,972 They are probed with purified proteins,930,973−975 cells,976 or living pathogens (Leishmania parasites).977 Binding events on these arrays are detected by fluorescence975,976,978,979 or SPRi.906,973,977,980 These arrays are prepared with full-length GAGs930 or with natural or synthetic oligosaccharides of defined sequence or length.979,981 Seeberger’s group has developed a general method for the automated assembly of glycans. It is carried out on a solid support with a

10.2. Interaction Networks of Glycosaminoglycans

The GAG−protein interactions can occur only with GAG fragments of defined sizes, which are integrated into the networks in order to reflect the physiological and pathological remodeling of the ECM, and to determine how it rewires the GAG interaction networks. These issues will be addressed below. GAGs interact with numerous partners and building their interaction networks is useful: (i) to investigate the crosstalk between the ECM and the intracellular or nuclear compartments, (ii) to identify the molecular connections between physio-pathological processes regulated by GAGs, and (iii) to AV

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

and 30 PGs.910 The syndecan and integrin interactomes are confined in small spaces, and their analysis gives insights into the composition of cell-matrix adhesion and their regulation.993 Two major roles of decorin maintenance on cell structure and outside-in signaling have been highlighted from its network analysis.217 In the global GAG interactome, one-third of GAGbinding proteins are involved in the regulation of cell growth and cell communication, whereas others contribute to ECM assembly and architecture.910

select protein−GAG interactions specifically associated with a pathological situation, which might be targeted by inhibitors disrupting these interactions.986,987 Hep/HS-protein interaction networks have been investigated in angiogenesis (the angiogenesis glycomic interactome)953 and at the endothelial cell surface.910 A strategy has been implemented to build protein−GAG interaction networks by combining experimental data collected by SPRi, SPR, and BLI, manual curation of the literature and interaction data stored in MatrixDB database (http://matrixdb.univ-lyon1.fr/)914,930 and in other databases of the International Molecular Exchange Consortium.988 The analysis of a global interactome comprising CS, DS, HA, Hep, and HS-binding proteins has shown that a number of GAGbinding proteins are able to bind to several GAGs, which results in a highly connected GAG−protein interaction network. All the proteins binding to DS interact with other GAGs, whereas Hep has the highest number of protein partners that bind only to it, but this may be a bias due to the fact that it is the most frequently studied GAG.930 A systems biology approach has been used to build the Hep/HS interactome, a protein−protein interaction network comprised of 435 Hep/HS binding-proteins identified by literature curation and chromatography of rat liver homogenate on immobilized Hep followed by mass spectrometry.919 Other protein−protein interaction networks of GAG-binding proteins have been built for individual Hep-binding proteins (e.g., antithrombin and FGF-1),918 Hep-binding proteins in disease (e.g., pancreatic diseases989 and periodontitis990), and in a cell (e.g., E. coli).985 GAG-binding protein interactomes (Figure 20) are analyzed with bioinformatics tools. The over-representation (enrichment) of Gene Ontology terms, domain families, and signaling pathways can be performed with the functional enrichment analysis tool FunRich (www.funrich.org),991 used for the analysis of the global GAG interactome,930 the Database for Annotation, Visualization and Integrated Discovery (DAVID, https://david.ncifcrf.gov/),992 used to analyze the Hep/HS interactome, or one of the Cytoscape apps for network analysis (http://apps.cytoscape.org/apps/with_tag/networkanalysis). Cytoscape is “an open source software platform for visualizing complex networks and integrating these with any type of attribute data” (http://www.cytoscape.org/).917 Topological parameters of networks (e.g., degree, diameter, shortest path length, clustering coefficient) can be calculated with the Cytoscape plugin Network Analyzer as for the Hep/HS interaction network.919 The above analyses are necessary to extract biological significance from the interaction networks. The analysis of the interactome of E. coli GAG-binding proteins has shown that Hep- and HS-interacting proteins play a role in Gly, Ser, and Thr metabolism and that the Hepbinding protein YcbS is a virulence factor.985

10.4. Integration of Structural, Kinetics, Affinity, and Expression Data in GAG/PG Interactomes

Tissue-specific interaction subnetworks of GAGs or PGs can be extracted from global networks by integrating transcriptomic and/or quantitative proteomic data. They are useful to understand how GAGs/PGs work in a concerted fashion in vivo and to decipher the molecular mechanisms underlying their biological roles in specific tissues and cells. Moreover, the integration of structural data in GAG/PG interactomes (e.g., protein binding sites and GAG features involved in protein interactions such as the number, position, and type of sulfate groups) will allow the building of 3D networks, the discrimination of competitive and noncompetitive interactions, the characterization of the glycocodes specifically regulating physiological and pathological processes, and the design of inhibitors disrupting specific GAG−protein interactions for therapeutic purpose. Kinetic parameters (association and dissociation rates) and the equilibrium dissociation constant KD reflecting the affinity can be integrated in the interaction networks to rank interactions. This has been done for the Hep-protein interaction network formed at the endothelial cell surface by endostatin, an antiangiogenic fragment of collagen type XVIII,910 and for the global GAG−protein interactome.930 The dissociation rate of GAG−protein complexes gives an estimation of their stability and discriminates transient interactions from stable ones. Thus, it would be interesting to determine how other major matrix components interact with their specific cell surface receptors through the GAG chains in the pericellular space and to what extent this relationship, in turn, contributes to the organization of surface PGs.

11. CONCLUSIONS AND PERSPECTIVES ECM is a 3D highly dynamic structural network. PGs, a major family of ECM molecules found in all cell types and tissues, contribute to the control of numerous normal and pathological processes. PGs present high structural complexity and heterogeneity. The major classes of PGs can be defined by their distribution, their sequence and domain organization homologies, and their functions in the ECM. Post-translational modifications of the PG protein core and secondary interactions with ECM enzymes also influence the PG biological roles. In the case of cell surface PGs, the functional diversity can be enhanced by the process of proteolytic ectodomain shedding, which converts membrane bound coreceptors into soluble paracrine effector molecules. GAGs are the building blocks of PGs that significantly influence their chemical and functional properties. The size of the GAG and its inherent sulfation patterns collectively form a glycocode deciphered by GAG-binding proteins and provide specific regulatory roles for each GAG protein core combination and defined structural motifs within GAGs that form the basis of rational drug design.

10.3. Interaction Networks of Proteoglycan Classes

The interaction networks of PG classes built by collecting interaction data stored in databases and those reported in the literature will be introduced in this section. Different layouts of these networks, integrating molecular functions of PGs and the biological processes they are involved in, are needed to understand if and how proteoglycan classes regulate biological processes in a concerted manner and to determine specific molecular mechanisms associated with each PG class. Interaction networks have been built for a single PG (i.e., decorin and serglycin),29,217 a family of PGs (syndecans),993 AW

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

book edition Extracellular Matrix: Pathobiology & Signaling and has received numerous awards. He serves as associate editor, academic editor, and editorial board member (EBM) in several journals (served also as EBM in J. Biol. Chem.), and he is currently the President of the Hellenic Society of Biochemistry & Molecular Biology and coordinator of the Matrix Biology section. His current research interests are focused on matrix pathobiology, cell signaling, molecular targeting, structural analysis of carbohydrates, preclinical evaluation of drugs at cell level and evaluation of cytotoxicity. More focus is given to proteoglycans, glycosaminoglycans, metalloproteinases, and acidic glycoproteins and especially to their structure−function relationship, implication in tissue organization, pathogenesis, and progression of various disorders, such as cancer and atherosclerosis.

HA, a nonsulfated GAG, is of remarkable biological importance. The size of HA plays a role in inflammation and cell behavior throughout the specific HA receptors. Control of the HA synthesis and epigenetics of its metabolism and biotechnological developments in the field are rapidly emerging areas of research and applications. Given that GAGs and PGs interact with many proteins (∼450 for Hep/HS), it is important to build their global interaction networks and tissue-specific interaction subnetworks in order to understand how they work in a concerted fashion in vivo and to decipher the molecular mechanisms underlying GAG/PG functions. Moreover, the integration of structural data in these interactomes (e.g., protein binding sites and GAG features involved in protein interactions such as the number and type of sulfate groups) is useful to build 3D networks, to identify the glycocodes regulating specific physiological and pathological processes, to discriminate competitive and noncompetitive interactions, and to select GAG−protein interactions as therapeutic targets. In recent years, several studies have been conducted in epigenetic level in order to better understand the biosynthetic mechanisms of PGs and the intracellular singaling cascades that they mediate. However, more in-depth studies are required to unravel the regulatory aspects of these mechanisms. The discovery of exosome composition revealed several major matrix components as their cargo, including the enzyme heparanase, that needs further investigation in order to investigate the regulatory mechanisms upon exosome biogenesis as well as in their paracrine interactions and biological functions. Future studies focusing on the clarification of exosome functions and interactions may contribute in the design and development of specific vehicles that inhibit their transport in several diseases or decoy receptors that recognize exosomes in order to avoid their attachment on the cell surface. At the biotechnological level, the controlled release of matrix macromolecules, such as endorepellin for autophagy, the inhibition of onco-miRNAs, and the combinational secretion of biomolecules in target tissues will improve the efficacy of novel diagnostic tools and disease progression. In conclusion, structure function relationships in terms of GAG/ PG synthesis, degradation, cell signaling, epigenetics, docking studies, and modeling as well as interacting networks (interactome) in relation to disease development and progression will be at the forefront of developing new diagnostic tools and potential pharmacological agents for disease targeting.

Zoi Piperigkou received her Ph.D. in Biochemistry, Cellular and Molecular Biology, under the supervision of Prof. Nikos K. Karamanos, at the University of Patras (Greece) in 2018. Her research interests have been focused on extracellular matrix, proteoglycans, glycosaminoglycans, epigenetics, and cancer cell pathobiology. She has also been trained in the implication of epigenetics, focusing on miRNAs in cancer progression, in Prof. Martin Götte’s laboratory in Münster (Germany), and she participates in competitive European Projects in collaboration with Foundation for Research and Technology-Hellas (FORTH)/Institute of Chemical Engineering Sciences (ICE-HT). Achilleas D. Theocharis is an Associate Professor of Biochemistry & Molecular Biology at the University of Patras. He obtained his Ph.D. in Biochemistry at the University of Patras in 2000. Following a postdoctoral training at the Karolinska Institute, Stockholm, Sweden, he was appointed as a faculty member at the Department of Chemistry, University of Patras, in 2003. His research interests are focused on the areas of matrix pathobiochemistry, cell signaling, and molecular targeting. He is editorial board member Matrix Biology, and he is coauthor in more than 90 publications in peer review international journals. Hideto Watanabe graduated from Kanazawa University School of Medicine (Japan) in 1985 (M.D.) and from Graduate School Kanazawa University (Japan) in 1989 (Ph.D.). During that time, he was trained in Professor Yutaka Nagai’s lab, working on collagens, collagenases, and leukocyte elastases. Then, he worked on matrix metalloproteinases with Professor Yasunori Okada. In 1992, he joined Dr. Yoshi Yamada’s lab at National Institutes of Health (U.S.A.) and studied on aggrecan and cartilage link protein, especially focusing on their gene knockout mice. In 2000, he joined Professor Koji Kimata’s lab and, since then, has been studying the role of versican in development and diseases, and synthesis of chondroitin sulfate synthesis. Since 2007, he is a Professor of Biochemistry at Aichi Medical University (Japan). He served as an editorial board member of several journals including J. Biol. Chem. and was President of the Japan Matrix Club. Since April in 2018, he has been the president of the Japanese Society for Matrix Biology and Medicine (JSMBM). Although his background is pathology, he has abundant experience of molecular biology, biochemistry, carbohydrate chemistry, and genetics.

AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected]. ORCID

Nikos K. Karamanos: 0000-0003-3618-0288 Notes

The authors declare no competing financial interest.

Marco Franchi is currently a researcher at the Department for Life Quality Studies, University of Bologna (Italy). He received his M.D. at the University of Bologna (Italy) in 1983 and his Ph.D. in Biomedical Technologies at the University of Bologna (Italy) in 1989. His research interests include the ultrastructure of collagen and collagen-proteoglycans interaction in fibrous connective tissues, tendons, and ligaments. His recent research interests are focused on cancer cell ultrastructure and functional collagen arrangement of

Biographies Nikos K. Karamanos is a Professor of Biochemistry and Organic Biochemical Analysis since 2003 in the University of Patras (Greece). He received his Ph.D. in Biochemistry from University of Patras and carried out pre- and postdoctoral research work at Karolinska Institute (School of Medicine, Sweden). He is the editor of the international AX

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

characterization of structure-interaction-function relationships of the extracellular matrix and of extracellular interaction networks with a special focus on collagen, matricryptin, and glycosaminoglycan networks as well as on the role of intrinsic disorder and mutations in rewiring these networks in diseases. She has been President of the French Society for Matrix Biology. She serves as associate editor of Matrix Biology, and she is currently the Vice-President of the International Society for Matrix Biology.

extracellular matrix in 3D cultures as microenvironment and barrier of cancer cells invasion and metastasis at the scanning electron microscope. Stéphanie Baud obtained her Ph.D. in Physics from the University of Besançon (France) in 2004 and then was a Postdoctoral Fellow at the Hospital for Sick Children in the Molecular Structure & Function department (Toronto, Canada). She joined the CNRS UMR 7369, at University of Reims Champagne Ardenne (URCA), in 2007 as an Associate Professor. During the last years, her research interests have been focused on deciphering at the atomic and molecular level (through molecular modeling and simulations) the processes governing the structure/dynamics/function relationships of extracellular matrix major macromolecules.

Ralph D. Sanderson received his Ph.D. degree in Cell Biology in 1986 from the University of Alabama at Birmingham followed by a postdoctoral fellowship at Stanford University. He joined the faculty in Pathology at the University of Arkansas for Medical Sciences in 1989 and held the Drs. Mae and Anderson Nettleship Chair in Oncologic Pathology from 2002−2006. He joined the UAB Department of Pathology in 2006 and currently is the UAB Endowed Professor in Cancer Pathobiology. His work is focused on the heparan sulfate proteoglycan syndecan-1 and the heparan sulfate degrading enzyme heparanase and their role in regulating tumor progression. Recently, his lab has focused on development and testing of inhibitors of heparanase and their potential as anticancer drugs and on the impact of syndecan-1 and heparanase on the biogenesis and function of tumor-derived exosomes.

Stéphane Brézillon is a senior researcher at CNRS Institute, UMR 7369, at University of Reims Champagne Ardenne (URCA), Reims, France. He received a joint Ph.D. in Cell Biology at URCA, Reims, France, and the Faculty of Medicine, Hannover, Germany in 1997. He was a Postdoctoral Fellow at the University of Saint-Andrews, Scotland, UK, until 1999. His recent interests include the inhibitory effect of lumican, a small leucine-rich proteoglycan, in tumor progression and the characterization of glycosaminoglycans by biophotonic approaches.

Thomas Neill obtained his Ph.D. in Cellular and Developmental Biology in 2014 from Thomas Jefferson University (PA, USA), where he focused on elucidating the roles of decorin mediated-suppression of tumorigenesis and angiogenesis. He then began a postdoctoral fellowship with Dr. Renato V. Iozzo at Thomas Jefferson University to focus on the roles of decorin and endorepellin in activating procatabolic signaling pathways for autophagic and mitophagic induction.

Martin Götte is a Professor for Medical Biochemistry at the University of Münster, Germany. He obtained his Ph.D. in Biochemistry at the University of Göttingen/Max-Planck-Institute for Biophysical Chemistry in 1997. Following a postdoctoral training at the Department of Cell Biology of Harvard Medical School, and a group leader position at the Institute of Physiological Chemistry of Münster University in 2000, he holds a tenured position as Head of Research in the Department of Gynecology and Obstetrics since 2003. His main areas of research are the role of Syndecans, Decorin, and GAG biosynthetic enzymes in cancer and inflammation. He is spokesman of the University in the Federal State network for stem cell research and chairman of the board for the reproduction section of the German Society for Endocrinology. He is editorial board member of four SCIlisted journals and has authored more than 120 publications with >5900 citations (H-index = 35).

Renato V. Iozzo is the Gonzalo E. Aponte Endowed Chair Professor of Pathology, Anatomy and Cell Biology at Thomas Jefferson University (PA, USA). Dr. Iozzo has received many awards including the Benjamin Castleman Award from the International Academy of Pathology, the Junior Faculty Research Award from the American Cancer Society, the Burlington Resources Foundation Faculty Achievement Award, the Faculty Research Award from the American Cancer Society, the Research Career Award from Thomas Jefferson University, and the Senior Investigator Award from the American Society for Matrix Biology. He was the chair of the Gordon Research Conference on Proteoglycans and President of both the International and American Societies for Matrix Biology. He has received Honorary Degrees from Semmelveis University (2011) and University of Patras (2016). Dr. Iozzo is Editor-in-Chief of Matrix Biology. His research focuses on proteoglycans and their roles in cancer, tumor angiogenesis, and autophagy. He has published >370 articles, reaching >40,000 citations and a h-index of 106.

Alberto Passi is a Professor of Biochemistry at University of Insubria (Varese, Italy) since 2008. He received his M.D. at the University of Pavia (Italy) in 1987 and his Ph.D. in Biochemistry and Clinical Chemistry at University of Pavia and Genoa (Italy) in 1992. He visited several laboratories worldwide and spent two years in the Bioengineering Laboratory in the Lerner Institute, Cleveland Clinic Foundation (USA). His main interest is the metabolism of hyaluronan and other glycosaminoglycans in mammals. He is actively involved in the biotechnological applications of hyaluronan in regenerative medicine being inventor of three international patents.

ACKNOWLEDGMENTS N.K.K., Z.P., A.D.T., and M.G. acknowledge support from EU Horizon 2020 project RISE-2014, action no. 645756 “GLYCANC − Matrix glycans as multifunctional pathogenesis factors and therapeutic targets in cancer”. Z.P. acknowledges support from State Scholarships Foundation (Greece) Grant no. 5003404 through the Operational Program “Human Resources Development, Education and Lifelong Learning”, of NSRF 2014-2020 by the cofunding of the European Social Fund. S.R.B. acknowledges grants of the European Commission (PSIMEx, contract FP7-HEALTH-2007-223411), from t h e F o n d a t i o n p o u r l a R e c h e r c h e M é d i c a l e (DBI20141231336), and from the Institut Français de Bioinformatique (ANR-11-INS-0013, Glycomatrix project),

Davide Vigetti is an Associate Professor of Biochemistry at the School of Medicine of University of Insubria, Varese (Italy). He received his Ph.D. in Evolutionary and Developmental Biology from University of Milan (Italy). His research interests include the cell biology and pathology of connective tissue focusing on vascular and tumors glycobiology. Specific interests include molecular aspect of the regulation of hyaluronic acid and proteoglycans synthesis investigating epigenetics and protein post-translational modifications. Sylvie Ricard-Blum is a Professor in Biochemistry at the University of Lyon (France). She received a Diplôme d’Etudes Approfondies (DEA) in Biochemistry and a Ph.D. in Biochemistry at the University of Lyon (France). She joined successively the Pasteur Institute in Lyon, the Institute for Structural Biology in Grenoble (France), and the University of Lyon 1 in 2004. Her research interests lie in the AY

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

GLCE GlcNAc GPCRs GPI HA HAPLNs HARE HAS HC HCC H/D Hep HepV HGF Hh HMW HPSE HS HS2ST HYALs IdoA IFN-γ IGF-IR Ihh IL ITC KD KO KS LDL LLC LMW LRRs LTBPs LYVE lncRNA mAb MAMs MAPK MBL MEFs MET miRNA MK MMPs MST MT-MMP 4-MU MVs NACs NAT NDST NF-κB

as well as S.D. Vallet (Lyon, France) for his help in designing Figure 20. R.S. acknowledges support from NIH grants CA138340 and CA211752 and the United States-Israel Binational Science Foundation. R.V.I. acknowledges support from NIH Grants CA39481 and CA47282. R.V.I. and T.N. acknowledge all members of Iozzo’s laboratory for valuable discussions and suggestions. S.B. acknowledges Pr. M. Dauchez, Dr. N. Belloy, and K. Karamanou for useful discussions and Dr. N. Belloy for designing the high-quality Figure 7.

ABBREVIATIONS AA amino acids ADAMs a disintegrin and metalloproteinases ADAMTs ADAMs with thrombospondin motifs Acan Aggrecan AVC atrioventricular canal B4GALT7 B1−4-galactosyltransferase-4 glucosaminyltransferase-7 B3GNT7 B1−3-N-acetylglucosaminyltransferase-7 BDNF brain-derived neurotrophic factor BLI biolayer interferometry CCL chemokine (C−C motif) ligand ceRNA competing endogenous RNA CHPF chondroitin polymerizing factor CHST carbohydrate sulfotransferases CHSY chondroitin synthases CNS central nervous system CRP complement regulatory protein CS chondroitin sulfate CSPG chondroitin sulfate proteoglycan CST chondroitin sulfotransferase CTGF connective tissue growth factor CTF C-terminal fragment 3D three-dimensional Da Dalton Dcn decorin DDRs discoidin domain receptors Dhh desert hedgehog DS dermatan sulfate DSTP dispirotripiperazine derivative Δψm mitochondrial membrane potential EAE experimental autoimmune encephalomyelitis ECM extracellular matrix EGF epidermal growth factor EGFR epidermal growth factor receptor Egr early growth response EMT epithelial-to-mesenchymal transition ER estrogen receptor ERK extracellular signal-regulated kinase ES endostatin ESCRT endosomal sorting complex required for transport EVs extracellular vesicles EXTL exostosin-like glycosyltransferase FAK focal adhesion kinase FGF fibroblast growth factor GAGs glycosaminoglycans Gal galactose GalNAc N-acetylgalactosamine GALT galactosyltransferase GAT glucuronosyltransferase GFP green fluorescence protein GlcA glucuronic acid

NLS NMR NO NICD PAPS PDGF PDZ PGs AZ

D-glucuronyl C5-epimerase N-acetylglucosamine G-protein-coupled heptahelical receptors glycosylphosphatidylinositol hyaluronan hyaluronan and PG link proteins hyaluronan receptor for endocytosis hyaluronan synthase heavy chain hepatocellular carcinoma hydrogen/deuterium heparin fragment of the α1(v) collagen chain hepatocyte growth factor hedgehog high molecular weight heparanase heparan sulfate HS 2-O-sulfotransferase hyaluronidases iduronic acid interferon-γ insulin-like growth factor receptor I Indian hedgehog interleukin isothermal titration calorimetry dissociation constant knockout keratan sulfate low density lipoprotein Lewis lung carcinoma low molecular weight leucine-rich repeats latent TGF-β-binding proteins lymphatic vessel endothelial HA receptor long noncoding RNA monoclonal antibody mitochondrial-associated membranes mitogen-activated protein kinase mannose binding lectin mouse embryonic fibroblasts mesenchymal-to-epithelial transition microRNA midkine metalloproteinases microscale thermophoresis membrane-type MMP 4-methylumbelliferone microvesicles neutrophil activating chemokines natural antisense transcript N-deacetylase/N-sulfotransferases nuclear factor kappa-light-chain-enhancer of activated B cells nuclear localization sequences nuclear magnetic resonance nitrate ions notch intracellular domain 3-phospho-5-adenylyl sulfate platelet-derived growth factor postsynaptic density 95/disc-large/zona occludens-1 proteoglycans DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews PI3K PKC PLGA PT PTM RPTPs QCM-D RHAMM RhoGDI ROS RTK SA SEM Shh Sulfs SLRPs SPR TEM TGF TLRs TRPC TSG UA2ST UDP UGDH Vcan VEGF VEGFR VSD WT Xyl

Review

(12) Sonnenschein, C.; Soto, A. M. Carcinogenesis Explained within the Context of a Theory of Organisms. Prog. Biophys. Mol. Biol. 2016, 122, 70−76. (13) Maffini, M. V.; Soto, A. M.; Calabro, J. M.; Ucci, A. A.; Sonnenschein, C. The Stroma as a Crucial Target in Rat Mammary Gland Carcinogenesis. J. Cell Sci. 2004, 117, 1495−1502. (14) Barcellos-Hoff, M. H.; Ravani, S. A. Irradiated Mammary Gland Stroma Promotes the Expression of Tumorigenic Potential by Unirradiated Epithelial Cells. Cancer Res. 2000, 60, 1254−1260. (15) Guerra, L.; Odorisio, T.; Zambruno, G.; Castiglia, D. Stromal Microenvironment in Type Vii Collagen-Deficient Skin: The Ground for Squamous Cell Carcinoma Development. Matrix Biol. 2017, 63, 1−10. (16) Iozzo, R. V.; Gubbiotti, M. A. Extracellular Matrix: The Driving Force of Mammalian Diseases. Matrix Biol. 2018, DOI: 10.1016/ j.matbio.2018.03.023. (17) Iozzo, R. V.; Schaefer, L. Proteoglycan Form and Function: A Comprehensive Nomenclature of Proteoglycans. Matrix Biol. 2015, 42, 11−55. (18) 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. (19) 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, Rev. Cancer 2015, 1855, 276−300. (20) Afratis, N. A.; Nikitovic, D.; Multhaupt, H. A.; Theocharis, A. D.; Couchman, J. R.; Karamanos, N. K. Syndecans - Key Regulators of Cell Signaling and Biological Functions. FEBS J. 2017, 284, 27−41. (21) Theocharis, A. D.; Gialeli, C.; Bouris, P.; Giannopoulou, E.; Skandalis, S. S.; Aletras, A. J.; Iozzo, R. V.; Karamanos, N. K. CellMatrix Interactions: Focus on Proteoglycan-Proteinase Interplay and Pharmacological Targeting in Cancer. FEBS J. 2014, 281, 5023−5042. (22) 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. (23) Soares da Costa, D.; Reis, R. L.; Pashkuleva, I. Sulfation of Glycosaminoglycans and Its Implications in Human Health and Disorders. Annu. Rev. Biomed. Eng. 2017, 19, 1−26. (24) Karousou, E.; Misra, S.; Ghatak, S.; Dobra, K.; Gotte, M.; Vigetti, D.; Passi, A.; Karamanos, N. K.; Skandalis, S. S. Roles and Targeting of the Has/Hyaluronan/Cd44 Molecular System in Cancer. Matrix Biol. 2017, 59, 3−22. (25) Almond, A. Multiscale Modeling of Glycosaminoglycan Structure and Dynamics: Current Methods and Challenges. Curr. Opin. Struct. Biol. 2018, 50, 58−64. (26) Rnjak-Kovacina, J.; Tang, F.; Whitelock, J. M.; Lord, M. S. Glycosaminoglycan and Proteoglycan-Based Biomaterials: Current Trends and Future Perspectives. Adv. Healthcare Mater. 2018, 7, No. e1701042. (27) Mende, M.; Bednarek, C.; Wawryszyn, M.; Sauter, P.; Biskup, M. B.; Schepers, U.; Brase, S. Chemical Synthesis of Glycosaminoglycans. Chem. Rev. 2016, 116, 8193−8255. (28) Iozzo, R. V. The Biology of the Small Leucine-Rich Proteoglycans. Functional Network of Interactive Proteins. J. Biol. Chem. 1999, 274, 18843−18846. (29) Korpetinou, A.; Skandalis, S. S.; Labropoulou, V. T.; Smirlaki, G.; Noulas, A.; Karamanos, N. K.; Theocharis, A. D. Serglycin: At the Crossroad of Inflammation and Malignancy. Front. Oncol. 2014, 3, 327. (30) Morawski, M.; Bruckner, G.; Arendt, T.; Matthews, R. T. Aggrecan: Beyond Cartilage and into the Brain. Int. J. Biochem. Cell Biol. 2012, 44, 690−693. (31) Wight, T. N. Provisional Matrix: A Role for Versican and Hyaluronan. Matrix Biol. 2017, 60−61, 38−56.

phosphatidylinositide 3-kinase protein kinase C poly lactic-co-glycolic acid pleiotrophin post-translational modification protein tyrosine phosphatase receptor quartz crystal microbalance with dissipation receptor for HA-mediated motility RhoGTP dissociation inhibitor reactive oxygen species receptor tyrosine kinase streptavidin scanning electron microscopy sonic hedgehog endosulfatases small leucine-rich PGs surface plasmon resonance transmission electron microscope transforming growth factor toll-like receptors transient receptor potential canonical channel TNF-stimulated gene uronosyl 2-sulfo- transferase uridine diphosphate UDP-glucose dehydrogenase Versican vascular endothelial growth factor vascular endothelial growth factor receptor ventricular septal defect wild-type xylose

REFERENCES (1) Frantz, C.; Stewart, K. M.; Weaver, V. M. The Extracellular Matrix at a Glance. J. Cell Sci. 2010, 123, 4195−4200. (2) Theocharis, A. D.; Skandalis, S. S.; Gialeli, C.; Karamanos, N. K. Extracellular Matrix Structure. Adv. Drug Delivery Rev. 2016, 97, 4− 27. (3) Ringer, P.; Colo, G.; Fassler, R.; Grashoff, C. Sensing the Mechano-Chemical Properties of the Extracellular Matrix. Matrix Biol. 2017, 64, 6−16. (4) Pozzi, A.; Yurchenco, P. D.; Iozzo, R. V. The Nature and Biology of Basement Membranes. Matrix Biol. 2017, 57−58, 1−11. (5) Multhaupt, H. A.; Leitinger, B.; Gullberg, D.; Couchman, J. R. Extracellular Matrix Component Signaling in Cancer. Adv. Drug Delivery Rev. 2016, 97, 28−40. (6) Theocharis, A. D.; Karamanos, N. K. Proteoglycans Remodeling in Cancer: Underlying Molecular Mechanisms. Matrix Biol. 2017, DOI: 10.1016/j.matbio.2017.10.008. (7) Gialeli, C.; Theocharis, A. D.; Karamanos, N. K. Roles of Matrix Metalloproteinases in Cancer Progression and Their Pharmacological Targeting. FEBS J. 2011, 278, 16−27. (8) Sanderson, R. D.; Elkin, M.; Rapraeger, A. C.; Ilan, N.; Vlodavsky, I. Heparanase Regulation of Cancer, Autophagy and Inflammation: New Mechanisms and Targets for Therapy. FEBS J. 2017, 284, 42−55. (9) Chester, D.; Brown, A. C. The Role of Biophysical Properties of Provisional Matrix Proteins in Wound Repair. Matrix Biol. 2017, 60− 61, 124−140. (10) Piperigkou, Z.; Gotte, M.; Theocharis, A. D.; Karamanos, N. K. Insights into the Key Roles of Epigenetics in Matrix MacromoleculesAssociated Wound Healing. Adv. Drug Delivery Rev. 2018, 129, 16− 36. (11) Bonnans, C.; Chou, J.; Werb, Z. Remodelling the Extracellular Matrix in Development and Disease. Nat. Rev. Mol. Cell Biol. 2014, 15, 786−801. BA

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(32) Theocharis, A. D. Versican in Health and Disease. Connect. Tissue Res. 2008, 49, 230−234. (33) Du, W. W.; Fang, L.; Yang, W.; Sheng, W.; Zhang, Y.; Seth, A.; Yang, B. B.; Yee, A. J. The Role of Versican G3 Domain in Regulating Breast Cancer Cell Motility Including Effects on Osteoblast Cell Growth and Differentiation in Vitro - Evaluation Towards Understanding Breast Cancer Cell Bone Metastasis. BMC Cancer 2012, 12, 341. (34) Du, W. W.; Yang, B. B.; Shatseva, T. A.; Yang, B. L.; Deng, Z.; Shan, S. W.; Lee, D. Y.; Seth, A.; Yee, A. J. Versican G3 Promotes Mouse Mammary Tumor Cell Growth, Migration, and Metastasis by Influencing Egf Receptor Signaling. PLoS One 2010, 5, e13828. (35) Buraschi, S.; Neill, T.; Iozzo, R. V. Decorin Is a Devouring Proteoglycan: Remodeling of Intracellular Catabolism Via Autophagy and Mitophagy. Matrix Biol. 2017, DOI: 10.1016/j.matbio.2017.10.005. (36) Schaefer, L.; Tredup, C.; Gubbiotti, M. A.; Iozzo, R. V. Proteoglycan Neofunctions: Regulation of Inflammation and Autophagy in Cancer Biology. FEBS J. 2017, 284, 10−26. (37) Gubbiotti, M. A.; Neill, T.; Iozzo, R. V. A Current View of Perlecan in Physiology and Pathology: A Mosaic of Functions. Matrix Biol. 2017, 57−58, 285−298. (38) Mundhenke, C.; Meyer, K.; Drew, S.; Friedl, A. Heparan Sulfate Proteoglycans as Regulators of Fibroblast Growth Factor-2 Receptor Binding in Breast Carcinomas. Am. J. Pathol. 2002, 160, 185−194. (39) Wang, H.; Jin, H.; Rapraeger, A. C. Syndecan-1 and Syndecan-4 Capture Epidermal Growth Factor Receptor Family Members and the Alpha3beta1 Integrin Via Binding Sites in Their Ectodomains: Novel Synstatins Prevent Kinase Capture and Inhibit Alpha6beta4-IntegrinDependent Epithelial Cell Motility. J. Biol. Chem. 2015, 290, 26103− 26113. (40) Wang, H.; Jin, H.; Beauvais, D. M.; Rapraeger, A. C. Cytoplasmic Domain Interactions of Syndecan-1 and Syndecan-4 with Alpha6beta4 Integrin Mediate Human Epidermal Growth Factor Receptor (Her1 and Her2)-Dependent Motility and Survival. J. Biol. Chem. 2014, 289, 30318−30332. (41) Huang, G.; Ge, G.; Izzi, V.; Greenspan, D. S. Alpha3 Chains of Type V Collagen Regulate Breast Tumour Growth Via Glypican-1. Nat. Commun. 2017, 8, 14351. (42) Filmus, J.; Capurro, M. Glypican-3: A Marker and a Therapeutic Target in Hepatocellular Carcinoma. FEBS J. 2013, 280, 2471−2476. (43) Han, S.; Ma, X.; Zhao, Y.; Zhao, H.; Batista, A.; Zhou, S.; Zhou, X.; Yang, Y.; Wang, T.; Bi, J.; et al. Identification of Glypican-3 as a Potential Metastasis Suppressor Gene in Gastric Cancer. Oncotarget 2016, 7, 44406−44416. (44) Buchanan, C.; Stigliano, I.; Garay-Malpartida, H. M.; Rodrigues Gomes, L.; Puricelli, L.; Sogayar, M. C.; Bal de Kier Joffe, E.; Peters, M. G. Glypican-3 Reexpression Regulates Apoptosis in Murine Adenocarcinoma Mammary Cells Modulating Pi3k/Akt and P38mapk Signaling Pathways. Breast Cancer Res. Treat. 2010, 119, 559−574. (45) Stigliano, I.; Puricelli, L.; Filmus, J.; Sogayar, M. C.; Bal de Kier Joffe, E.; Peters, M. G. Glypican-3 Regulates Migration, Adhesion and Actin Cytoskeleton Organization in Mammary Tumor Cells through Wnt Signaling Modulation. Breast Cancer Res. Treat. 2009, 114, 251− 262. (46) Castillo, L. F.; Tascon, R.; Lago Huvelle, M. A.; Novack, G.; Llorens, M. C.; dos Santos, A. F.; Shortrede, J.; Cabanillas, A. M.; Bal de Kier Joffe, E.; Labriola, L.; et al. Glypican-3 Induces a Mesenchymal to Epithelial Transition in Human Breast Cancer Cells. Oncotarget 2016, 7, 60133−60154. (47) Capurro, M. I.; Xu, P.; Shi, W.; Li, F.; Jia, A.; Filmus, J. Glypican-3 Inhibits Hedgehog Signaling During Development by Competing with Patched for Hedgehog Binding. Dev. Cell 2008, 14, 700−711. (48) Cattaruzza, S.; Nicolosi, P. A.; Braghetta, P.; Pazzaglia, L.; Benassi, M. S.; Picci, P.; Lacrima, K.; Zanocco, D.; Rizzo, E.; Stallcup, W. B.; et al. Ng2/Cspg4-Collagen Type Vi Interplays Putatively

Involved in the Microenvironmental Control of Tumour Engraftment and Local Expansion. J. Mol. Cell Biol. 2013, 5, 176−193. (49) Makagiansar, I. T.; Williams, S.; Mustelin, T.; Stallcup, W. B. Differential Phosphorylation of Ng2 Proteoglycan by Erk and Pkcalpha Helps Balance Cell Proliferation and Migration. J. Cell Biol. 2007, 178, 155−165. (50) Makagiansar, I. T.; Williams, S.; Dahlin-Huppe, K.; Fukushi, J.; Mustelin, T.; Stallcup, W. B. Phosphorylation of Ng2 Proteoglycan by Protein Kinase C-Alpha Regulates Polarized Membrane Distribution and Cell Motility. J. Biol. Chem. 2004, 279, 55262−55270. (51) Iida, J.; Wilhelmson, K. L.; Ng, J.; Lee, P.; Morrison, C.; Tam, E.; Overall, C. M.; McCarthy, J. B. Cell Surface Chondroitin Sulfate Glycosaminoglycan in Melanoma: Role in the Activation of ProMmp-2 (Pro-Gelatinase a). Biochem. J. 2007, 403, 553−563. (52) Lindahl, U.; Couchman, J.; Kimata, K.; Esko, J. D. In Essentials of Glycobiology; Varki, A., Cummings, R. D., Esko, J. D., Stanley, P., Hart, G. W., Aebi, M., Darvill, A. G., Kinoshita, T., Packer, N. H.et al., Eds. Cold Spring Harbor: NY, 2015, DOI: 10.1101/glycobiology.3e.017. (53) Esko, J. D.; Kimata, K.; Lindahl, U. In Essentials of Glycobiology, 2nd ed.; Varki, A., Cummings, R. D., Esko, J. D., Freeze, H. H., Stanley, P., Bertozzi, C. R., Etzler, M. E., Eds.; Cold Spring Harbor Laboratory Press: New York, 2009. (54) Lamari, F.; Theocharis, A.; Hjerpe, A.; Karamanos, N. K. Ultrasensitive Capillary Electrophoresis of Sulfated Disaccharides in Chondroitin/Dermatan Sulfates by Laser-Induced Fluorescence after Derivatization with 2-Aminoacridone. J. Chromatogr., Biomed. Appl. 1999, 730, 129−133. (55) Karamanos, N. K.; Vanky, P.; Syrokou, A.; Hjerpe, A. Identity of Dermatan and Chondroitin Sequences in Dermatan Sulfate Chains Determined by Using Fragmentation with Chondroitinases and IonPair High-Performance Liquid Chromatography. Anal. Biochem. 1995, 225, 220−230. (56) Karamanos, N. K.; Syrokou, A.; Vanky, P.; Nurminen, M.; Hjerpe, A. Determination of 24 Variously Sulfated Galactosaminoglycan- and Hyaluronan-Derived Disaccharides by High-Performance Liquid Chromatography. Anal. Biochem. 1994, 221, 189−199. (57) Sugahara, K.; Mikami, T. Chondroitin/Dermatan Sulfate in the Central Nervous System. Curr. Opin. Struct. Biol. 2007, 17, 536−545. (58) Mizumoto, S.; Yamada, S.; Sugahara, K. Molecular Interactions between Chondroitin-Dermatan Sulfate and Growth Factors/ Receptors/Matrix Proteins. Curr. Opin. Struct. Biol. 2015, 34, 35−42. (59) Malavaki, C.; Mizumoto, S.; Karamanos, N.; Sugahara, K. Recent Advances in the Structural Study of Functional Chondroitin Sulfate and Dermatan Sulfate in Health and Disease. Connect. Tissue Res. 2008, 49, 133−139. (60) Weiss, R. J.; Esko, J. D.; Tor, Y. Targeting Heparin and Heparan Sulfate Protein Interactions. Org. Biomol. Chem. 2017, 15, 5656−5668. (61) Karamanos, N. K.; Hjerpe, A. A Survey of Methodological Challenges for Glycosaminoglycan/Proteoglycan Analysis and Structural Characterization by Capillary Electrophoresis. Electrophoresis 1998, 19, 2561−2571. (62) Karamanos, N. K.; Vanky, P.; Tzanakakis, G. N.; Hjerpe, A. High Performance Capillary Electrophoresis Method to Characterize Heparin and Heparan Sulfate Disaccharides. Electrophoresis 1996, 17, 391−395. (63) Militsopoulou, M.; Lamari, F. N.; Hjerpe, A.; Karamanos, N. K. Determination of Twelve Heparin- and Heparan Sulfate-Derived Disaccharides as 2-Aminoacridone Derivatives by Capillary Zone Electrophoresis Using Ultraviolet and Laser-Induced Fluorescence Detection. Electrophoresis 2002, 23, 1104−1109. (64) Quantock, A. J.; Young, R. D.; Akama, T. O. Structural and Biochemical Aspects of Keratan Sulphate in the Cornea. Cell. Mol. Life Sci. 2010, 67, 891−906. (65) Littlechild, S. L.; Young, R. D.; Caterson, B.; Yoshida, H.; Yamazaki, M.; Sakimura, K.; Quantock, A. J.; Akama, T. O. Keratan Sulfate Phenotype in the Beta-1,3-N-Acetylglucosaminyltransferase-7BB

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Null Mouse Cornea. Invest. Ophthalmol. Visual Sci. 2018, 59, 1641− 1651. (66) Skandalis, S. S.; Gialeli, C.; Theocharis, A. D.; Karamanos, N. K. Advances and Advantages of Nanomedicine in the Pharmacological Targeting of Hyaluronan-Cd44 Interactions and Signaling in Cancer. Adv. Cancer Res. 2014, 123, 277−317. (67) Esko, J. D.; J, H. P.; Linhardt, R. J. In Essentials of Glycobiology; Varki, A., Cummings, R. D., Esko, J. D., Stanley, P., Hart, G. W., Aebi, M.; Darvill, A. G.; Kinoshita, T.; Packer, N. H.et al., Eds. Cold Spring Harbor: NY, 2015, DOI: 10.1101/glycobiology.3e.038. (68) Kjellen, L.; Lindahl, U. Specificity of GlycosaminoglycanProtein Interactions. Curr. Opin. Struct. Biol. 2018, 50, 101−108. (69) Xu, D.; Esko, J. D. Demystifying Heparan Sulfate-Protein Interactions. Annu. Rev. Biochem. 2014, 83, 129−157. (70) Pomin, V. H. Biological Findings from the Recent Nmr-Based Studies of Glycosaminoglycan-Protein Interactions. Glycobiology 2014, 24, 991−1003. (71) Mizumoto, S.; Fongmoon, D.; Sugahara, K. Interaction of Chondroitin Sulfate and Dermatan Sulfate from Various Biological Sources with Heparin-Binding Growth Factors and Cytokines. Glycoconjugate J. 2013, 30, 619−632. (72) Li, W.; Johnson, D. J.; Esmon, C. T.; Huntington, J. A. Structure of the Antithrombin-Thrombin-Heparin Ternary Complex Reveals the Antithrombotic Mechanism of Heparin. Nat. Struct. Mol. Biol. 2004, 11, 857−862. (73) Brunetti, J.; Depau, L.; Falciani, C.; Gentile, M.; Mandarini, E.; Riolo, G.; Lupetti, P.; Pini, A.; Bracci, L. Insights into the Role of Sulfated Glycans in Cancer Cell Adhesion and Migration through Use of Branched Peptide Probe. Sci. Rep. 2016, 6, 27174. (74) Chikama, S.; Iida, S.; Inoue, M.; Kawagoe, N.; Tomiyasu, K.; Matsuoka, K.; Noda, S.; Takazono, I. Role of Heparan Sulfate Proteoglycan (Syndecan-1) on the Renal Epithelial Cells During Calcium Oxalate Monohydrate Crystal Attachment. Kurume Med. J. 2002, 49, 201−210. (75) Rajarathnam, K.; Sepuru, K. M.; Joseph, P. R. B.; Sawant, K. V.; Brown, A. J. Glycosaminoglycan Interactions Fine-Tune ChemokineMediated Neutrophil Trafficking: Structural Insights and Molecular Mechanisms. J. Histochem. Cytochem. 2018, 66, 229−239. (76) Djerbal, L.; Lortat-Jacob, H.; Kwok, J. Chondroitin Sulfates and Their Binding Molecules in the Central Nervous System. Glycoconjugate J. 2017, 34, 363−376. (77) Deepa, S. S.; Carulli, D.; Galtrey, C.; Rhodes, K.; Fukuda, J.; Mikami, T.; Sugahara, K.; Fawcett, J. W. Composition of Perineuronal Net Extracellular Matrix in Rat Brain: A Different Disaccharide Composition for the Net-Associated Proteoglycans. J. Biol. Chem. 2006, 281, 17789−17800. (78) Sugahara, K.; Mikami, T.; Uyama, T.; Mizuguchi, S.; Nomura, K.; Kitagawa, H. Recent Advances in the Structural Biology of Chondroitin Sulfate and Dermatan Sulfate. Curr. Opin. Struct. Biol. 2003, 13, 612−620. (79) Sugiura, N.; Shioiri, T.; Chiba, M.; Sato, T.; Narimatsu, H.; Kimata, K.; Watanabe, H. Construction of a Chondroitin Sulfate Library with Defined Structures and Analysis of Molecular Interactions. J. Biol. Chem. 2012, 287, 43390−43400. (80) Mizumoto, S.; Takahashi, J.; Sugahara, K. Receptor for Advanced Glycation End Products (Rage) Functions as Receptor for Specific Sulfated Glycosaminoglycans, and Anti-Rage Antibody or Sulfated Glycosaminoglycans Delivered in Vivo Inhibit Pulmonary Metastasis of Tumor Cells. J. Biol. Chem. 2012, 287, 18985−18994. (81) Liang, W. G.; Triandafillou, C. G.; Huang, T. Y.; Zulueta, M. M.; Banerjee, S.; Dinner, A. R.; Hung, S. C.; Tang, W. J. Structural Basis for Oligomerization and Glycosaminoglycan Binding of Ccl5 and Ccl3. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 5000−5005. (82) Richter, R. P.; Baranova, N. S.; Day, A. J.; Kwok, J. C. Glycosaminoglycans in Extracellular Matrix Organisation: Are Concepts from Soft Matter Physics Key to Understanding the Formation of Perineuronal Nets? Curr. Opin. Struct. Biol. 2018, 50, 65−74.

(83) Blundell, C. D.; Almond, A.; Mahoney, D. J.; DeAngelis, P. L.; Campbell, I. D.; Day, A. J. Towards a Structure for a Tsg6.Hyaluronan Complex by Modeling and Nmr Spectroscopy: Insights into Other Members of the Link Module Superfamily. J. Biol. Chem. 2005, 280, 18189−18201. (84) Lawrance, W.; Banerji, S.; Day, A. J.; Bhattacharjee, S.; Jackson, D. G. Binding of Hyaluronan to the Native Lymphatic Vessel Endothelial Receptor Lyve-1 Is Critically Dependent on Receptor Clustering and Hyaluronan Organization. J. Biol. Chem. 2016, 291, 8014−8030. (85) Yingsung, W.; Zhuo, L.; Morgelin, M.; Yoneda, M.; Kida, D.; Watanabe, H.; Ishiguro, N.; Iwata, H.; Kimata, K. Molecular Heterogeneity of the Shap-Hyaluronan Complex. Isolation and Characterization of the Complex in Synovial Fluid from Patients with Rheumatoid Arthritis. J. Biol. Chem. 2003, 278, 32710−32718. (86) Boraldi, F.; Croce, M. A.; Quaglino, D.; Sammarco, R.; Carnevali, E.; Tiozzo, R.; Pasquali-Ronchetti, I. Cell-Matrix Interactions of in Vitro Human Skin Fibroblasts Upon Addition of Hyaluronan. Tissue Cell 2003, 35, 37−45. (87) Jacoboni, I.; Valdre, U.; Mori, G.; Quaglino, D., Jr.; PasqualiRonchetti, I. Hyaluronic Acid by Atomic Force Microscopy. J. Struct. Biol. 1999, 126, 52−58. (88) Dickinson, L. E.; Gerecht, S. Micropatterned Surfaces to Study Hyaluronic Acid Interactions with Cancer Cells. J. Visualized Exp. 2010, 46, DOI: 10.3791/2413. (89) Fioravanti, A.; Cantarini, L.; Chellini, F.; Manca, D.; Paccagnini, E.; Marcolongo, R.; Collodel, G. Effect of Hyaluronic Acid (Mw 500−730 Kda) on Proteoglycan and Nitric Oxide Production in Human Osteoarthritic Chondrocyte Cultures Exposed to Hydrostatic Pressure. Osteoarthritis Cartilage 2005, 13, 688−696. (90) Glowacki, J.; Trepman, E.; Folkman, J. Cell Shape and Phenotypic Expression in Chondrocytes. Exp. Biol. Med. 1983, 172, 93−98. (91) Turner, N. J.; Kielty, C. M.; Walker, M. G.; Canfield, A. E. A Novel Hyaluronan-Based Biomaterial (Hyaff-11) as a Scaffold for Endothelial Cells in Tissue Engineered Vascular Grafts. Biomaterials 2004, 25, 5955−5964. (92) Pasquinelli, G.; Orrico, C.; Foroni, L.; Bonafe, F.; Carboni, M.; Guarnieri, C.; Raimondo, S.; Penna, C.; Geuna, S.; Pagliaro, P.; et al. Mesenchymal Stem Cell Interaction with a Non-Woven HyaluronanBased Scaffold Suitable for Tissue Repair. J. Anat. 2008, 213, 520− 530. (93) Karamanou, K.; Franchi, M.; Piperigkou, Z.; Perreau, C.; Maquart, F. X.; Vynios, D. H.; Brezillon, S. Lumican Effectively Regulates the Estrogen Receptors-Associated Functional Properties of Breast Cancer Cells, Expression of Matrix Effectors and Epithelial-toMesenchymal Transition. Sci. Rep. 2017, 7, 45138. (94) Piperigkou, Z.; Karamanou, K.; Afratis, N. A.; Bouris, P.; Gialeli, C.; Belmiro, C. L.; Pavao, M. S.; Vynios, D. H.; Tsatsakis, A. M. Biochemical and Toxicological Evaluation of Nano-Heparins in Cell Functional Properties, Proteasome Activation and Expression of Key Matrix Molecules. Toxicol. Lett. 2016, 240, 32−42. (95) Afratis, N. A.; Karamanou, K.; Piperigkou, Z.; Vynios, D. H.; Theocharis, A. D. The Role of Heparins and Nano-Heparins as Therapeutic Tool in Breast Cancer. Glycoconjugate J. 2017, 34, 299− 307. (96) Pavao, M. S. Glycosaminoglycans Analogs from Marine Invertebrates: Structure, Biological Effects, and Potential as New Therapeutics. Front. Cell. Infect. Microbiol. 2014, 4, 123. (97) Gerotziafas, G. T.; Papageorgiou, C.; Hatmi, M.; Samama, M. M.; Elalamy, I. Clinical Studies with Anticoagulants to Improve Survival in Cancer Patients. Pathophysiol. Haemostasis Thromb. 2009, 36, 204−211. (98) Miller, G. J.; Hansen, S. U.; Barath, M.; Johannessen, C.; Blanch, E. W.; Jayson, G. C.; Gardiner, J. M. Synthesis of a HeparinRelated Glcn-Idoa Sulfation-Site Variable Disaccharide Library and Analysis by Raman and Roa Spectroscopy. Carbohydr. Res. 2014, 400, 44−53. BC

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Toxicity in Xenografts of Human Breast Cancer. J. Controlled Release 2010, 146, 212−218. (115) Yu, D.; Li, W.; Zhang, Y.; Zhang, B. Anti-Tumor Efficiency of Paclitaxel and DNA When Co-Delivered by Ph Responsive Ligand Modified Nanocarriers for Breast Cancer Treatment. Biomed. Pharmacother. 2016, 83, 1428−1435. (116) Bachar, G.; Cohen, K.; Hod, R.; Feinmesser, R.; Mizrachi, A.; Shpitzer, T.; Katz, O.; Peer, D. Hyaluronan-Grafted Particle Clusters Loaded with Mitomycin C as Selective Nanovectors for Primary Head and Neck Cancers. Biomaterials 2011, 32, 4840−4848. (117) Zaga, T.; Makris, D.; Tsilioni, I.; Kiropoulos, T.; Oikonomidi, S.; Damianos, A.; Gourgoulianis, K. I. Hyaluronic Acid Levels Are Increased in Complicated Parapneumonic Pleural Effusions. Monaldi Arch. Chest Dis. 2011, 75, 167−171. (118) Cortes-Dericks, L.; Schmid, R. A. Cd44 and Its Ligand Hyaluronan as Potential Biomarkers in Malignant Pleural Mesothelioma: Evidence and Perspectives. Respir. Res. 2017, 18, 58. (119) Creaney, J.; Dick, I. M.; Segal, A.; Musk, A. W.; Robinson, B. W. Pleural Effusion Hyaluronic Acid as a Prognostic Marker in Pleural Malignant Mesothelioma. Lung Cancer 2013, 82, 491−498. (120) Dobra, K.; Andang, M.; Syrokou, A.; Karamanos, N. K.; Hjerpe, A. Differentiation of Mesothelioma Cells Is Influenced by the Expression of Proteoglycans. Exp. Cell Res. 2000, 258, 12−22. (121) Thylen, A.; Hjerpe, A.; Martensson, G. Hyaluronan Content in Pleural Fluid as a Prognostic Factor in Patients with Malignant Pleural Mesothelioma. Cancer 2001, 92, 1224−1230. (122) Karamanos, N. K.; Hjerpe, A. High-Performance Capillary Electrophoretic Analysis of Hyaluronan in Effusions from Human Malignant Mesothelioma. J. Chromatogr., Biomed. Appl. 1997, 697, 277−281. (123) Cheng, X. B.; Sato, N.; Kohi, S.; Yamaguchi, K. Prognostic Impact of Hyaluronan and Its Regulators in Pancreatic Ductal Adenocarcinoma. PLoS One 2013, 8, e80765. (124) Chen, C. P.; Hung, W.; Lin, S. H. Effectiveness of Hyaluronic Acid for Treating Diabetic Foot: A Systematic Review and MetaAnalysis. Dermatol. Ther. 2014, 27, 331−336. (125) McHutchison, J. G.; Blatt, L. M.; de Medina, M.; Craig, J. R.; Conrad, A.; Schiff, E. R.; Tong, M. J. Measurement of Serum Hyaluronic Acid in Patients with Chronic Hepatitis C and Its Relationship to Liver Histology. Consensus Interferon Study Group. J. Gastroenterol. Hepatol. 2000, 15, 945−951. (126) Zheng, M.; Cai, W. M.; Zhao, J. K.; Zhu, S. M.; Liu, R. H. Determination of Serum Levels of Ykl-40 and Hyaluronic Acid in Patients with Hepatic Fibrosis Due to Schistosomiasis Japonica and Appraisal of Their Clinical Value. Acta Trop. 2005, 96, 148−152. (127) Fried, M. W.; Duncan, A.; Soroka, S.; Connaghan, D. G.; Farrand, A.; Peter, J.; Strauss, R. M.; Boyer, T. D.; McDonald, G. B. Serum Hyaluronic Acid in Patients with Veno-Occlusive Disease Following Bone Marrow Transplantation. Bone Marrow Transplant. 2001, 27, 635−639. (128) Emlen, W.; Niebur, J.; Flanders, G.; Rutledge, J. Measurement of Serum Hyaluronic Acid in Patients with Rheumatoid Arthritis: Correlation with Disease Activity. J. Rheumatol. 1996, 23, 974−978. (129) Chua, J. S.; Kuberan, B. Synthetic Xylosides: Probing the Glycosaminoglycan Biosynthetic Machinery for Biomedical Applications. Acc. Chem. Res. 2017, 50, 2693−2705. (130) Eriksson, B. I.; Quinlan, D. J. Oral Anticoagulants in Development: Focus on Thromboprophylaxis in Patients Undergoing Orthopaedic Surgery. Drugs 2006, 66, 1411−1429. (131) Eikelboom, J. W.; Weitz, J. I. New Oral Anticoagulants for Thromboprophylaxis in Patients Having Hip or Knee Arthroplasty. BMJ 2011, 342, c7270. (132) Persson, A.; Tykesson, E.; Westergren-Thorsson, G.; Malmstrom, A.; Ellervik, U.; Mani, K. Xyloside-Primed Chondroitin Sulfate/Dermatan Sulfate from Breast Carcinoma Cells with a Defined Disaccharide Composition Has Cytotoxic Effects in Vitro. J. Biol. Chem. 2016, 291, 14871−14882. (133) Christianson, H. C.; Svensson, K. J.; van Kuppevelt, T. H.; Li, J. P.; Belting, M. Cancer Cell Exosomes Depend on Cell-Surface

(99) Ferro, V.; Dredge, K.; Liu, L.; Hammond, E.; Bytheway, I.; Li, C.; Johnstone, K.; Karoli, T.; Davis, K.; Copeman, E.; et al. Pi-88 and Novel Heparan Sulfate Mimetics Inhibit Angiogenesis. Semin. Thromb. Hemostasis 2007, 33, 557−568. (100) 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. (101) Galli, M.; Chatterjee, M.; Grasso, M.; Specchia, G.; Magen, H.; Einsele, H.; Celeghini, I.; Barbieri, P.; Paoletti, D.; Pace, S.et al. Phase I Study of the Heparanase Inhibitor Roneparstat: An Innovative Approach for Multiple Myeloma Therapy. Haematologica 2018, . (102) Pumphrey, C. Y.; Theus, A. M.; Li, S.; Parrish, R. S.; Sanderson, R. D. Neoglycans, Carbodiimide-Modified Glycosaminoglycans: A New Class of Anticancer Agents That Inhibit Cancer Cell Proliferation and Induce Apoptosis. Cancer Res. 2002, 62, 3722− 3728. (103) Yang, H.; Liu, S.; Cai, H.; Wan, L.; Li, S.; Li, Y.; Cheng, J.; Lu, X. Chondroitin Sulfate as a Molecular Portal That Preferentially Mediates the Apoptotic Killing of Tumor Cells by PenetratinDirected Mitochondria-Disrupting Peptides. J. Biol. Chem. 2010, 285, 25666−25676. (104) Lensen, J. F.; van der Vlag, J.; Versteeg, E. M.; Wetzels, J. F.; van den Heuvel, L. P.; Berden, J. H.; van Kuppevelt, T. H.; Rops, A. L. Differential Expression of Specific Dermatan Sulfate Domains in Renal Pathology. PLoS One 2015, 10, e0134946. (105) Mashima, R.; Sakai, E.; Tanaka, M.; Kosuga, M.; Okuyama, T. The Levels of Urinary Glycosaminoglycans of Patients with Attenuated and Severe Type of Mucopolysaccharidosis Ii Determined by Liquid Chromatography-Tandem Mass Spectrometry. Mol. Genet. Metab. Rep. 2016, 7, 87−91. (106) Rochette, A.; Boufaied, N.; Scarlata, E.; Hamel, L.; Brimo, F.; Whitaker, H. C.; Ramos-Montoya, A.; Neal, D. E.; Dragomir, A.; Aprikian, A.; et al. Asporin Is a Stromally Expressed Marker Associated with Prostate Cancer Progression. Br. J. Cancer 2017, 116, 775−784. (107) Karalis, T. T.; Heldin, P.; Vynios, D. H.; Neill, T.; Buraschi, S.; Iozzo, R. V.; Karamanos, N. K.; Skandalis, S. S. Tumor-Suppressive Functions of 4-Mu on Breast Cancer Cells of Different Er Status: Regulation of Hyaluronan/Has2/Cd44 and Specific Matrix Effectors. Matrix Biol. 2018, DOI: DOI: 10.1016/j.matbio.2018.04.007 10.1016/j.matbio.2018.04.007. (108) Yates, T. J.; Lopez, L. E.; Lokeshwar, S. D.; Ortiz, N.; Kallifatidis, G.; Jordan, A.; Hoye, K.; Altman, N.; Lokeshwar, V. B. Dietary Supplement 4-Methylumbelliferone: An Effective Chemopreventive and Therapeutic Agent for Prostate Cancer. J. Natl. Cancer Inst. 2015, 107, DOI: 10.1093/jnci/djv085. (109) Misra, S.; Heldin, P.; Hascall, V. C.; Karamanos, N. K.; Skandalis, S. S.; Markwald, R. R.; Ghatak, S. Hyaluronan-Cd44 Interactions as Potential Targets for Cancer Therapy. FEBS J. 2011, 278, 1429−1443. (110) Pouyani, T.; Prestwich, G. D. Functionalized Derivatives of Hyaluronic Acid Oligosaccharides: Drug Carriers and Novel Biomaterials. Bioconjugate Chem. 1994, 5, 339−347. (111) Xu, X.; Jha, A. K.; Harrington, D. A.; Farach-Carson, M. C.; Jia, X. Hyaluronic Acid-Based Hydrogels: From a Natural Polysaccharide to Complex Networks. Soft Matter 2012, 8, 3280− 3294. (112) Luo, Y.; Ziebell, M. R.; Prestwich, G. D. A Hyaluronic AcidTaxol Antitumor Bioconjugate Targeted to Cancer Cells. Biomacromolecules 2000, 1, 208−218. (113) Tariq, M.; Alam, M. A.; Singh, A. T.; Panda, A. K.; Talegaonkar, S. Improved Oral Efficacy of Epirubicin through Polymeric Nanoparticles: Pharmacodynamic and Toxicological Investigations. Drug Delivery 2016, 23, 2990−2997. (114) Cai, S.; Thati, S.; Bagby, T. R.; Diab, H. M.; Davies, N. M.; Cohen, M. S.; Forrest, M. L. Localized Doxorubicin Chemotherapy with a Biopolymeric Nanocarrier Improves Survival and Reduces BD

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

particles with Antimicrobial Properties. Nanomedicine (London, U. K.) 2009, 4, 421−429. (154) Abdel-Mohsen, A. M.; Hrdina, R.; Burgert, L.; Abdel-Rahman, R. M.; Hasova, M.; Smejkalova, D.; Kolar, M.; Pekar, M.; Aly, A. S. Antibacterial Activity and Cell Viability of Hyaluronan Fiber with Silver Nanoparticles. Carbohydr. Polym. 2013, 92, 1177−1187. (155) Zhang, X.; Yao, M.; Chen, M.; Li, L.; Dong, C.; Hou, Y.; Zhao, H.; Jia, B.; Wang, F. Hyaluronic Acid-Coated Silver Nanoparticles as a Nanoplatform for in Vivo Imaging Applications. ACS Appl. Mater. Interfaces 2016, 8, 25650−25653. (156) Young, J. J.; Cheng, K. M.; Young, Y. A.; Chen, X. A.; Chen, Y. H.; Chang, T. Y.; Yen, H. J.; Chen, C. C. Chondroitin SulfateStabilized Silver Nanoparticles: Improved Synthesis and Their Catalytic, Antimicrobial, and Biocompatible Activities. Carbohydr. Res. 2018, 457, 14−24. (157) Varghese, O. P.; Liu, J.; Sundaram, K.; Hilborn, J.; Oommen, O. P. Chondroitin Sulfate Derived Theranostic Nanoparticles for Targeted Drug Delivery. Biomater. Sci. 2016, 4, 1310−1313. (158) Oommen, O. P.; Duehrkop, C.; Nilsson, B.; Hilborn, J.; Varghese, O. P. Multifunctional Hyaluronic Acid and Chondroitin Sulfate Nanoparticles: Impact of Glycosaminoglycan Presentation on Receptor Mediated Cellular Uptake and Immune Activation. ACS Appl. Mater. Interfaces 2016, 8, 20614−20624. (159) Coburn, J. M.; Gibson, M.; Monagle, S.; Patterson, Z.; Elisseeff, J. H. Bioinspired Nanofibers Support Chondrogenesis for Articular Cartilage Repair. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 10012−10017. (160) Harley, B. A.; Lynn, A. K.; Wissner-Gross, Z.; Bonfield, W.; Yannas, I. V.; Gibson, L. J. Design of a Multiphase Osteochondral Scaffold Iii: Fabrication of Layered Scaffolds with Continuous Interfaces. J. Biomed. Mater. Res., Part A 2010, 92, 1078−1093. (161) Lynn, A. K.; Best, S. M.; Cameron, R. E.; Harley, B. A.; Yannas, I. V.; Gibson, L. J.; Bonfield, W. Design of a Multiphase Osteochondral Scaffold. I. Control of Chemical Composition. J. Biomed. Mater. Res., Part A 2010, 92, 1057−1065. (162) Kim, B. K.; Kim, D.; Kwak, G.; Yhee, J. Y.; Kwon, I. C.; Kim, S. H.; Yeo, Y. Polyethylenimine-Dermatan Sulfate Complex, a Bioactive Biomaterial with Unique Toxicity to Cd146-Positive Cancer Cells. ACS Biomater. Sci. Eng. 2017, 3, 990−999. (163) Wen, Y.; Grondahl, L.; Gallego, M. R.; Jorgensen, L.; Mo̷ ller, E. H.; Nielsen, H. M. Delivery of Dermatan Sulfate from Polyelectrolyte Complex-Containing Alginate Composite Microspheres for Tissue Regeneration. Biomacromolecules 2012, 13, 905− 917. (164) Anthony, C.; Anna, P.; Lennart, R. In Proteoglycans; Iozzo, R. V., Ed.; Marcel Dekker, Inc: New York, 2000. (165) Watanabe, H.; Yamada, Y.; Kimata, K. Roles of Aggrecan, a Large Chondroitin Sulfate Proteoglycan, in Cartilage Structure and Function. J. Biochem. 1998, 124, 687−693. (166) Chang, Y.; Yanagishita, M.; Hascall, V. C.; Wight, T. N. Proteoglycans Synthesized by Smooth Muscle Cells Derived from Monkey (Macaca Nemestrina) Aorta. J. Biol. Chem. 1983, 258, 5679− 5688. (167) Kimata, K.; Oike, Y.; Tani, K.; Shinomura, T.; Yamagata, M.; Uritani, M.; Suzuki, S. A Large Chondroitin Sulfate Proteoglycan (PgM) Synthesized before Chondrogenesis in the Limb Bud of Chick Embryo. J. Biol. Chem. 1986, 261, 13517−13525. (168) Kang, J. S.; Oohashi, T.; Kawakami, Y.; Bekku, Y.; Izpisua Belmonte, J. C.; Ninomiya, Y. Characterization of Dermacan, a Novel Zebrafish Lectican Gene, Expressed in Dermal Bones. Mech. Dev. 2004, 121, 301−312. (169) Wight, T. N. Versican: A Versatile Extracellular Matrix Proteoglycan in Cell Biology. Curr. Opin. Cell Biol. 2002, 14, 617− 623. (170) Wight, T. N.; Kinsella, M. G.; Evanko, S. P.; Potter-Perigo, S.; Merrilees, M. J. Versican and the Regulation of Cell Phenotype in Disease. Biochim. Biophys. Acta, Gen. Subj. 2014, 1840, 2441−2451.

Heparan Sulfate Proteoglycans for Their Internalization and Functional Activity. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, 17380−17385. (134) Raman, K.; Ninomiya, M.; Nguyen, T. K.; Tsuzuki, Y.; Koketsu, M.; Kuberan, B. Novel Glycosaminoglycan Biosynthetic Inhibitors Affect Tumor-Associated Angiogenesis. Biochem. Biophys. Res. Commun. 2011, 404, 86−89. (135) Rolls, A.; Shechter, R.; London, A.; Segev, Y.; Jacob-Hirsch, J.; Amariglio, N.; Rechavi, G.; Schwartz, M. Two Faces of Chondroitin Sulfate Proteoglycan in Spinal Cord Repair: A Role in Microglia/ Macrophage Activation. PLoS Med. 2008, 5, e171. (136) Ferdous, Z.; Grande-Allen, K. J. Utility and Control of Proteoglycans in Tissue Engineering. Tissue Eng. 2007, 13, 1893− 1904. (137) Weyers, A.; Linhardt, R. J. Neoproteoglycans in Tissue Engineering. FEBS J. 2013, 280, 2511−2522. (138) Lutolf, M. P.; Hubbell, J. A. Synthetic Biomaterials as Instructive Extracellular Microenvironments for Morphogenesis in Tissue Engineering. Nat. Biotechnol. 2005, 23, 47−55. (139) Li, Y.; Wang, Y.; Huang, G.; Gao, J. Cooperativity Principles in Self-Assembled Nanomedicine. Chem. Rev. 2018, 118, 5359−5391. (140) Fubini, B.; Ghiazza, M.; Fenoglio, I. Physico-Chemical Features of Engineered Nanoparticles Relevant to Their Toxicity. Nanotoxicology 2010, 4, 347−363. (141) Davis, M. E.; Chen, Z. G.; Shin, D. M. Nanoparticle Therapeutics: An Emerging Treatment Modality for Cancer. Nat. Rev. Drug Discovery 2008, 7, 771−782. (142) Cavalcante, M. C.; Allodi, S.; Valente, A. P.; Straus, A. H.; Takahashi, H. K.; Mourao, P. A.; Pavao, M. S. Occurrence of Heparin in the Invertebrate Styela Plicata (Tunicata) Is Restricted to Cell Layers Facing the Outside Environment. An Ancient Role in Defense? J. Biol. Chem. 2000, 275, 36189−36186. (143) Santos, J. C.; Mesquita, J. M.; Belmiro, C. L.; da Silveira, C. B.; Viskov, C.; Mourier, P. A.; Pavao, M. S. Isolation and Characterization of a Heparin with Low Antithrombin Activity from the Body of Styela Plicata (Chordata-Tunicata). Distinct Effects on Venous and Arterial Models of Thrombosis. Thromb. Res. 2007, 121, 213−223. (144) Murugesan, S.; Park, T. J.; Yang, H.; Mousa, S.; Linhardt, R. J. Blood Compatible Carbon Nanotubes–Nano-Based Neoproteoglycans. Langmuir 2006, 22, 3461−3463. (145) Huang, H.; Yang, X. Synthesis of Polysaccharide-Stabilized Gold and Silver Nanoparticles: A Green Method. Carbohydr. Res. 2004, 339, 2627−2631. (146) Kemp, M. M.; Kumar, A.; Mousa, S.; Dyskin, E.; Yalcin, M.; Ajayan, P.; Linhardt, R. J.; Mousa, S. A. Gold and Silver Nanoparticles Conjugated with Heparin Derivative Possess Anti-Angiogenesis Properties. Nanotechnology 2009, 20, 455104. (147) Kemp, M. M.; Linhardt, R. J. Heparin-Based Nanoparticles. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2010, 2, 77−87. (148) Nurunnabi, M.; Khatun, Z.; Moon, W. C.; Lee, G.; Lee, Y. K. Heparin Based Nanoparticles for Cancer Targeting and Noninvasive Imaging. Quantitative Imaging in Medicine and Surgery 2012, 2, 219− 226. (149) Chung, Y. I.; Kim, S. K.; Lee, Y. K.; Park, S. J.; Cho, K. O.; Yuk, S. H.; Tae, G.; Kim, Y. H. Efficient Revascularization by Vegf Administration Via Heparin-Functionalized Nanoparticle-Fibrin Complex. J. Controlled Release 2010, 143, 282−289. (150) She, W.; Li, N.; Luo, K.; Guo, C.; Wang, G.; Geng, Y.; Gu, Z. Dendronized Heparin-Doxorubicin Conjugate Based Nanoparticle as Ph-Responsive Drug Delivery System for Cancer Therapy. Biomaterials 2013, 34, 2252−2264. (151) Yang, X.; Du, H.; Liu, J.; Zhai, G. Advanced Nanocarriers Based on Heparin and Its Derivatives for Cancer Management. Biomacromolecules 2015, 16, 423−436. (152) Tan, Q.; Tang, H.; Hu, J.; Hu, Y.; Zhou, X.; Tao, Y.; Wu, Z. Controlled Release of Chitosan/Heparin Nanoparticle-Delivered Vegf Enhances Regeneration of Decellularized Tissue-Engineered Scaffolds. Int. J. Nanomed. 2011, 6, 929−942. (153) Kemp, M. M.; Kumar, A.; Clement, D.; Ajayan, P.; Mousa, S.; Linhardt, R. J. Hyaluronan- and Heparin-Reduced Silver NanoBE

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Differentiation and Regulates Joint Morphogenesis. J. Biol. Chem. 2010, 285, 21114−21125. (189) Hatano, S.; Nagai, N.; Sugiura, N.; Tsuchimoto, J.; Isogai, Z.; Kimata, K.; Ota, A.; Karnan, S.; Hosokawa, Y.; Watanabe, H. Versican a-Subdomain Is Required for Its Adequate Function in Dermal Development. Connect. Tissue Res. 2018, 59, 178−190. (190) Kishimoto, J.; Ehama, R.; Wu, L.; Jiang, S.; Jiang, N.; Burgeson, R. E. Selective Activation of the Versican Promoter by Epithelial- Mesenchymal Interactions During Hair Follicle Development. Proc. Natl. Acad. Sci. U. S. A. 1999, 96, 7336−7341. (191) Soma, T.; Tajima, M.; Kishimoto, J. Hair Cycle-Specific Expression of Versican in Human Hair Follicles. J. Dermatol. Sci. 2005, 39, 147−154. (192) Jo, S. J.; Kim, J. Y.; Jang, S.; Choi, S. J.; Kim, K. H.; Kwon, O. Decrease of Versican Levels in the Follicular Dermal Papilla Is a Remarkable Aging-Associated Change of Human Hair Follicles. J. Dermatol. Sci. 2016, 84, 354−357. (193) Wagner, R. B.; Johnston, M. R. Middle Lobe Syndrome. Ann. Thorac. Surg. 1983, 35, 679−686. (194) Burin-des-Roziers, C.; Rothschild, P. R.; Layet, V.; Chen, J. M.; Ghiotti, T.; Leroux, C.; Cremers, F. P.; Brezin, A. P.; Valleix, S. Deletions Overlapping Vcan Exon 8 Are New Molecular Defects for Wagner Disease. Hum. Mutat. 2017, 38, 43−47. (195) Rothschild, P. R.; Audo, I.; Nedelec, B.; Ghiotti, T.; Brezin, A. P.; Monin, C.; Valleix, S. De Novo Splice Mutation in the Versican Gene in a Family with Wagner Syndrome. JAMA Ophthalmol. 2013, 131, 805−807. (196) Kloeckener-Gruissem, B.; Neidhardt, J.; Magyar, I.; Plauchu, H.; Zech, J. C.; Morle, L.; Palmer-Smith, S. M.; Macdonald, M. J.; Nas, V.; Fry, A. E.; et al. Novel Vcan Mutations and Evidence for Unbalanced Alternative Splicing in the Pathogenesis of Wagner Syndrome. Eur. J. Hum. Genet. 2013, 21, 352−356. (197) Zheng, P. S.; Wen, J.; Ang, L. C.; Sheng, W.; Viloria-Petit, A.; Wang, Y.; Wu, Y.; Kerbel, R. S.; Yang, B. B. Versican/Pg-M G3 Domain Promotes Tumor Growth and Angiogenesis. FASEB J. 2004, 18, 754−756. (198) Arslan, F.; Bosserhoff, A. K.; Nickl-Jockschat, T.; Doerfelt, A.; Bogdahn, U.; Hau, P. The Role of Versican Isoforms V0/V1 in Glioma Migration Mediated by Transforming Growth Factor-Beta2. Br. J. Cancer 2007, 96, 1560−1568. (199) Serra, M.; Miquel, L.; Domenzain, C.; Docampo, M. J.; Fabra, A.; Wight, T. N.; Bassols, A. V3 Versican Isoform Expression Alters the Phenotype of Melanoma Cells and Their Tumorigenic Potential. Int. J. Cancer 2005, 114, 879−886. (200) Hernandez, D.; Miquel-Serra, L.; Docampo, M. J.; MarcoRamell, A.; Cabrera, J.; Fabra, A.; Bassols, A. V3 Versican Isoform Alters the Behavior of Human Melanoma Cells by Interfering with Cd44/Erbb-Dependent Signaling. J. Biol. Chem. 2011, 286, 1475− 1485. (201) Ricciardelli, C.; Mayne, K.; Sykes, P. J.; Raymond, W. A.; McCaul, K.; Marshall, V. R.; Horsfall, D. J. Elevated Levels of Versican but Not Decorin Predict Disease Progression in Early-Stage Prostate Cancer. Clin. Cancer Res. 1998, 4, 963−971. (202) Ricciardelli, C.; Brooks, J. H.; Suwiwat, S.; Sakko, A. J.; Mayne, K.; Raymond, W. A.; Seshadri, R.; LeBaron, R. G.; Horsfall, D. J. Regulation of Stromal Versican Expression by Breast Cancer Cells and Importance to Relapse-Free Survival in Patients with NodeNegative Primary Breast Cancer. Clin. Cancer Res. 2002, 8, 1054− 1060. (203) Kodama, J.; Hasengaowa; Kusumoto, T.; Seki, N.; Matsuo, T.; Ojima, Y.; Nakamura, K.; Hongo, A.; Hiramatsu, Y. Prognostic Significance of Stromal Versican Expression in Human Endometrial Cancer. Ann. Oncol. 2006, 18, 269−274. (204) Said, N.; Sanchez-Carbayo, M.; Smith, S. C.; Theodorescu, D. Rhogdi2 Suppresses Lung Metastasis in Mice by Reducing Tumor Versican Expression and Macrophage Infiltration. J. Clin. Invest. 2012, 122, 1503−1518. (205) Shen, X. H.; Lin, W. R.; Xu, M. D.; Qi, P.; Dong, L.; Zhang, Q. Y.; Ni, S. J.; Weng, W. W.; Tan, C.; Huang, D.; et al. Prognostic

(171) Zimmermann, D. R.; Ruoslahti, E. Multiple Domains of the Large Fibroblast Proteoglycan, Versican. EMBO J. 1989, 8, 2975− 2981. (172) Shinomura, T.; Nishida, Y.; Ito, K.; Kimata, K. Cdna Cloning of Pg-M, a Large Chondroitin Sulfate Proteoglycan Expressed During Chondrogenesis in Chick Limb Buds. Alternative Spliced Multiforms of Pg-M and Their Relationships to Versican. J. Biol. Chem. 1993, 268, 14461−14469. (173) Milev, P.; Maurel, P.; Chiba, A.; Mevissen, M.; Popp, S.; Yamaguchi, Y.; Margolis, R. K.; Margolis, R. U. Differential Regulation of Expression of Hyaluronan-Binding Proteoglycans in Developing Brain: Aggrecan, Versican, Neurocan, and Brevican. Biochem. Biophys. Res. Commun. 1998, 247, 207−212. (174) Matsumoto, K.; Shionyu, M.; Go, M.; Shimizu, K.; Shinomura, T.; Kimata, K.; Watanabe, H. Distinct Interaction of Versican/Pg-M with Hyaluronan and Link Protein. J. Biol. Chem. 2003, 278, 41205−41212. (175) Watanabe, H.; Cheung, S. C.; Itano, N.; Kimata, K.; Yamada, Y. Identification of Hyaluronan-Binding Domains of Aggrecan. J. Biol. Chem. 1997, 272, 28057−28065. (176) Day, A. J.; Prestwich, G. D. Hyaluronan-Binding Proteins: Tying up the Giant. J. Biol. Chem. 2002, 277, 4585−4588. (177) Suwan, K.; Choocheep, K.; Hatano, S.; Kongtawelert, P.; Kimata, K.; Watanabe, H. Versican/Pg-M Assembles Hyaluronan into Extracellular Matrix and Inhibits Cd44-Mediated Signaling toward Premature Senescence in Embryonic Fibroblasts. J. Biol. Chem. 2009, 284, 8596−8604. (178) Merrilees, M. J.; Zuo, N.; Evanko, S. P.; Day, A. J.; Wight, T. N. G1 Domain of Versican Regulates Hyaluronan Organization and the Phenotype of Cultured Human Dermal Fibroblasts. J. Histochem. Cytochem. 2016, 64, 353−363. (179) Murasawa, Y.; Watanabe, K.; Yoneda, M.; Zako, M.; Kimata, K.; Sakai, L. Y.; Isogai, Z. Homotypic Versican G1 Domain Interactions Enhance Hyaluronan Incorporation into Fibrillin Microfibrils. J. Biol. Chem. 2013, 288, 29170−29181. (180) Spicer, A. P.; Joo, A.; Bowling, R. A., Jr. A Hyaluronan Binding Link Protein Gene Family Whose Members Are Physically Linked Adjacent to Chondroitin Sulfate Proteoglycan Core Protein Genes: The Missing Links. J. Biol. Chem. 2003, 278, 21083−21091. (181) Wirrig, E. E.; Snarr, B. S.; Chintalapudi, M. R.; O’Neal, J. L.; Phelps, A. L.; Barth, J. L.; Fresco, V. M.; Kern, C. B.; Mjaatvedt, C. H.; Toole, B. P.; et al. Cartilage Link Protein 1 (Crtl1), an Extracellular Matrix Component Playing an Important Role in Heart Development. Dev. Biol. 2007, 310, 291−303. (182) Mjaatvedt, C. H.; Yamamura, H.; Capehart, A. A.; Turner, D.; Markwald, R. R. The Cspg2 Gene, Disrupted in the Hdf Mutant, Is Required for Right Cardiac Chamber and Endocardial Cushion Formation. Dev. Biol. 1998, 202, 56−66. (183) Henderson, D. J.; Copp, A. J. Versican Expression Is Associated with Chamber Specification, Septation, and Valvulogenesis in the Developing Mouse Heart. Circ. Res. 1998, 83, 523−532. (184) Hatano, S.; Kimata, K.; Hiraiwa, N.; Kusakabe, M.; Isogai, Z.; Adachi, E.; Shinomura, T.; Watanabe, H. Versican/Pg-M Is Essential for Ventricular Septal Formation Subsequent to Cardiac Atrioventricular Cushion Development. Glycobiology 2012, 22, 1268−1277. (185) Shibata, S.; Fukada, K.; Imai, H.; Abe, T.; Yamashita, Y. In Situ Hybridization and Immunohistochemistry of Versican, Aggrecan and Link Protein, and Histochemistry of Hyaluronan in the Developing Mouse Limb Bud Cartilage. J. Anat. 2003, 203, 425−432. (186) Kamiya, N.; Watanabe, H.; Habuchi, H.; Takagi, H.; Shinomura, T.; Shimizu, K.; Kimata, K. Versican/Pg-M Regulates Chondrogenesis as an Extracellular Matrix Molecule Crucial for Mesenchymal Condensation. J. Biol. Chem. 2006, 281, 2390−2400. (187) Williams, D. R., Jr.; Presar, A. R.; Richmond, A. T.; Mjaatvedt, C. H.; Hoffman, S.; Capehart, A. A. Limb Chondrogenesis Is Compromised in the Versican Deficient Hdf Mouse. Biochem. Biophys. Res. Commun. 2005, 334, 960−966. (188) Choocheep, K.; Hatano, S.; Takagi, H.; Watanabe, H.; Kimata, K.; Kongtawelert, P.; Watanabe, H. Versican Facilitates Chondrocyte BF

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Significance of Versican Expression in Gastric Adenocarcinoma. Oncogenesis 2015, 4, e178. (206) Keire, P. A.; Bressler, S. L.; Lemire, J. M.; Edris, B.; Rubin, B. P.; Rahmani, M.; McManus, B. M.; van de Rijn, M.; Wight, T. N. A Role for Versican in the Development of Leiomyosarcoma. J. Biol. Chem. 2014, 289, 34089−34103. (207) Keire, P. A.; Bressler, S. L.; Mulvihill, E. R.; Starcher, B. C.; Kang, I.; Wight, T. N. Inhibition of Versican Expression by Sirna Facilitates Tropoelastin Synthesis and Elastic Fiber Formation by Human Sk-Lms-1 Leiomyosarcoma Smooth Muscle Cells in Vitro and in Vivo. Matrix Biol. 2016, 50, 67−81. (208) Fanhchaksai, K.; Okada, F.; Nagai, N.; Pothacharoen, P.; Kongtawelert, P.; Hatano, S.; Makino, S.; Nakamura, T.; Watanabe, H. Host Stromal Versican Is Essential for Cancer-Associated Fibroblast Function to Inhibit Cancer Growth. Int. J. Cancer 2016, 138, 630−641. (209) Asano, K.; Nelson, C. M.; Nandadasa, S.; Aramaki-Hattori, N.; Lindner, D. J.; Alban, T.; Inagaki, J.; Ohtsuki, T.; Oohashi, T.; Apte, S. S.; et al. Stromal Versican Regulates Tumor Growth by Promoting Angiogenesis. Sci. Rep. 2017, 7, 17225. (210) Kim, S.; Takahashi, H.; Lin, W. W.; Descargues, P.; Grivennikov, S.; Kim, Y.; Luo, J. L.; Karin, M. Carcinoma-Produced Factors Activate Myeloid Cells through Tlr2 to Stimulate Metastasis. Nature 2009, 457, 102−106. (211) Gao, D.; Vahdat, L. T.; Wong, S.; Chang, J. C.; Mittal, V. Microenvironmental Regulation of Epithelial-Mesenchymal Transitions in Cancer. Cancer Res. 2012, 72, 4883−4889. (212) Kalamajski, S.; Oldberg, A. The Role of Small Leucine-Rich Proteoglycans in Collagen Fibrillogenesis. Matrix Biol. 2010, 29, 248− 253. (213) Chen, S.; Young, M. F.; Chakravarti, S.; Birk, D. E. Interclass Small Leucine-Rich Repeat Proteoglycan Interactions Regulate Collagen Fibrillogenesis and Corneal Stromal Assembly. Matrix Biol. 2014, 35, 103−111. (214) 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. (215) Dunkman, A. A.; Buckley, M. R.; Mienaltowski, M. J.; Adams, S. M.; Thomas, S. J.; Satchell, L.; Kumar, A.; Pathmanathan, L.; Beason, D. P.; Iozzo, R. V.; et al. Decorin Expression Is Important for Age-Related Changes in Tendon Structure and Mechanical Properties. Matrix Biol. 2013, 32, 3−13. (216) Dunkman, A. A.; Buckley, M. R.; Mienaltowski, M. J.; Adams, S. M.; Thomas, S. J.; Kumar, A.; Beason, D. P.; Iozzo, R. V.; Birk, D. E.; Soslowsky, L. J. The Injury Response of Aged Tendons in the Absence of Biglycan and Decorin. Matrix Biol. 2014, 35, 232−238. (217) Gubbiotti, M. A.; Vallet, S. D.; Ricard-Blum, S.; Iozzo, R. V. Decorin Interacting Network: A Comprehensive Analysis of DecorinBinding Partners and Their Versatile Functions. Matrix Biol. 2016, 55, 7−21. (218) Merline, R.; Moreth, K.; Beckmann, J.; Nastase, M. V.; ZengBrouwers, J.; Tralhao, J. G.; Lemarchand, P.; Pfeilschifter, J.; Schaefer, R. M.; Iozzo, R. V.; et al. Signaling by the Matrix Proteoglycan Decorin Controls Inflammation and Cancer through Pdcd4 and Microrna-21. Sci. Signaling 2011, 4, ra75. (219) Frey, T.; Schroeder, N.; Manon-Jensen, T.; Iozzo, R. V.; Schaefer, L. Biological Interplay between Proteoglycans and Their Innate Immune Receptors in Inflammation. FEBS J. 2013, 280, 2165−2179. (220) Patel, S.; Santra, M.; McQuillan, D. J.; Iozzo, R. V.; Thomas, A. P. Decorin Activates the Epidermal Growth Factor Receptor and Elevates Cytosolic Ca2+ in A431 Cells. J. Biol. Chem. 1998, 273, 3121−3124. (221) Jarvelainen, H.; Puolakkainen, P.; Pakkanen, S.; Brown, E. L.; Hook, M.; Iozzo, R. V.; Sage, H.; Wight, T. N. A Role for Decorin in Cutaneous Wound Healing and Angiogenesis. Wound Rep. Reg. 2006, 14, 443−452.

(222) Baghy, K.; Iozzo, R. V.; Kovalszky, I. Decorin-Tgfb Axis in Hepatic Fibrosis and Cirrhosis. J. Histochem. Cytochem. 2012, 60, 262−268. (223) Muir, A. M.; Massoudi, D.; Nguyen, N.; Keene, D. R.; Lee, S. J.; Birk, D. E.; Davidson, J. M.; Marinkovich, M. P.; Greenspan, D. S. Bmp1-Like Proteinases Are Essential to the Structure and Wound Healing of Skin. Matrix Biol. 2016, 56, 114−131. (224) Nikolovska, K.; Renke, J. K.; Jungmann, O.; Grobe, K.; Iozzo, R. V.; Zamfir, A. D.; Seidler, D. G. A Decorin-Deficient Matrix Affects Skin Chondroitin/Dermatan Sulfate Levels and Keratinocyte Function. Matrix Biol. 2014, 35, 91−102. (225) Merline, R.; Lazaroski, S.; Babelova, A.; Tsalastra-Greul, W.; Pfeilschifter, J.; Schluter, K. D.; Gunther, A.; Iozzo, R. V.; Schaefer, R. M.; Schaefer, L. Decorin Deficiency in Diabetic Mice: Aggravation of Nephropathy Due to Overexpression of Profibrotic Factors, Enhanced Apoptosis and Mononuclear Cell Infiltration. J. Physiol. Pharmacol. 2009, 60, 5−13. (226) Wu, Z.; Horgan, C. E.; Carr, O.; Owens, R. T.; Iozzo, R. V.; Lechner, B. E. Biglycan and Decorin Differentially Regulate Signaling in the Fetal Membranes. Matrix Biol. 2014, 35, 266−275. (227) Bolton, K.; Segal, D.; McMillan, J.; Jowett, J.; Heilbronn, L.; Abberton, K.; Zimmet, P.; Chisholm, D.; Collier, G.; Walder, K. Decorin Is a Secreted Protein Associated with Obesity and Type 2 Diabetes. Int. J. Obes. 2008, 32, 1113−1121. (228) Marchica, C. L.; Pinelli, V.; Borges, M.; Zummer, J.; Narayanan, V.; Iozzo, R. V.; Ludwig, M. S. A Role for Decorin in a Murine Model of Allergen-Induced Asthma. Am. J. Physiol. Lung Cell Mol. Physiol. 2011, 300, 863−873. (229) Bocian, C.; Urbanowitz, A. K.; Owens, R. T.; Iozzo, R. V.; Gotte, M.; Seidler, D. G. Decorin Potentiates Interferon-Gamma Activity in a Model of Allergic Inflammation. J. Biol. Chem. 2013, 288, 12699−12711. (230) Seidler, D. G.; Mohamed, N. A.; Bocian, C.; Stadtmann, A.; Hermann, S.; Schafers, K.; Schafers, M.; Iozzo, R. V.; Zarbock, A.; Gotte, M. The Role for Decorin in Delayed-Type Hypersensitivity. J. Immunol. 2011, 187, 6108−6199. (231) 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. (232) 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. (233) Neill, T.; Jones, H. R.; Crane-Smith, Z.; Owens, R. T.; Schaefer, L.; Iozzo, R. V. Decorin Induces Rapid Secretion of Thrombospondin-1 in Basal Breast Carcinoma Cells Via Inhibition of Ras Homolog Gene Family, Member a/Rho-Associated Coiled-Coil Containing Protein Kinase 1. FEBS J. 2013, 280, 2353−2368. (234) Neill, T.; Schaefer, L.; Iozzo, R. V. Decorin, a Guardian from the Matrix. Am. J. Pathol. 2012, 181, 380−387. (235) Baghy, K.; Dezso, K.; Laszlo, V.; Fullar, A.; Peterfia, B.; Paku, S.; Nagy, P.; Schaff, Z.; Iozzo, R. V.; Kovalszky, I. Ablation of the Decorin Gene Enhances Experimental Hepatic Fibrosis and Impairs Hepatic Healing in Mice. Lab. Invest. 2011, 91, 439−451. (236) Horvath, Z.; Kovalszky, I.; Fullar, A.; Kiss, K.; Schaff, Z.; Iozzo, R. V.; Baghy, K. Decorin Deficiency Promotes Hepatic Carcinogenesis. Matrix Biol. 2014, 35, 194−205. (237) Duncan, M. B. Extracellular Matrix Transcriptome Dynamics in Hepatocellular Carcinoma. Matrix Biol. 2013, 32, 393−398. (238) Brandan, E.; Cabello-Verrugio, C.; Vial, C. Novel Regulatory Mechanisms for the Proteoglycans Decorin and Biglycan During Muscle Formation and Muscular Dystrophy. Matrix Biol. 2008, 27, 700−708. (239) Brandan, E.; Gutierrez, J. Role of Skeletal Muscle Proteoglycans During Myogenesis. Matrix Biol. 2013, 32, 289−297. (240) Weis, S. M.; Zimmerman, S. D.; Shah, M.; Covell, J. W.; Omens, J. H.; Ross, J., Jr.; Dalton, N.; Jones, Y.; Reed, C. C.; Iozzo, R. V.; et al. A Role for Decorin in the Remodeling of Myocardial Infarction. Matrix Biol. 2005, 24, 313−324. BG

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Engineered Alterations in Collagen and Decorin. J. Biomech. Eng. 2004, 126, 252−257. (260) Seidler, D. G.; Faiyaz-Ul-Haque, M.; Hansen, U.; Yip, G. W.; Zaidi, S. H.; Teebi, A. S.; Kiesel, L.; Gotte, M. Defective Glycosylation of Decorin and Biglycan, Altered Collagen Structure, and Abnormal Phenotype of the Skin Fibroblasts of an Ehlers-Danlos Syndrome Patient Carrying the Novel Arg270cys Substitution in Galactosyltransferase I (Beta4galt-7). J. Mol. Med. (Heidelberg, Ger.) 2006, 84, 583−594. (261) Moreth, K.; Iozzo, R. V.; Schaefer, L. Small Leucine-Rich Proteoglycans Orchestrate Receptor Crosstalk During Inflammation. Cell Cycle 2012, 11, 2084−2091. (262) Neill, T.; Schaefer, L.; Iozzo, R. V. An Oncosuppressive Role for Decorin. Mol. Cell. Oncol. 2015, 2, e975645. (263) Goldoni, S.; Iozzo, R. V. Tumor Microenvironment: Modulation by Decorin and Related Molecules Harboring LeucineRich Tandem Motifs. Int. J. Cancer 2008, 123, 2473−2479. (264) Iozzo, R. V.; Buraschi, S.; Genua, M.; Xu, S.-Q.; Solomides, C. C.; Peiper, S. C.; Gomella, L. G.; Owens, R. T.; Morrione, A. Decorin Antagonizes Igf Receptor I (Igf-Ir) Function by Interfering with Igf-Ir Activity and Attenuating Downstream Signaling. J. Biol. Chem. 2011, 286, 34712−34721. (265) Iozzo, R. V.; Chakrani, F.; Perrotti, D.; McQuillan, D. J.; Skorski, T.; Calabretta, B.; Eichstetter, I. Cooperative Action of Germline Mutations in Decorin and P53 Accelerates Lymphoma Tumorigenesis. Proc. Natl. Acad. Sci. U. S. A. 1999, 96, 3092−3097. (266) Bi, X.; Tong, C.; Dockendorff, A.; Bancroft, L.; Gallagher, L.; Guzman, G.; Iozzo, R. V.; Augenlicht, L. H.; Yang, W. Genetic Deficiency of Decorin Causes Intestinal Tumor Formation through Disruption of Intestinal Cell Maturation. Carcinogenesis 2008, 29, 1435−1440. (267) Bi, X.; Pohl, N. M.; Yang, G. R.; Gou, Y.; Guzman, G.; Kajdacsy-Balla, A.; Iozzo, R. V.; Yang, W.; Qian, Z. Decorin-Mediated Inhibition of Colorectal Cancer Growth and Migration Is Associated with E-Cadherin in Vitro and in Mice. Carcinogenesis 2012, 33, 326− 330. (268) Bi, X.; Xia, X.; Fan, D.; Mu, T.; Zhang, Q.; Iozzo, R. V.; Yang, W. Oncogenic Activin C Interacts with Decorin in Colorectal Cancer in Vivo and in Vitro. Mol. Carcinog. 2016, 55, 1786−1795. (269) Goldoni, S.; Owens, R. T.; McQuillan, D. J.; Shriver, Z.; Sasisekharan, R.; Birk, D. E.; Campbell, S.; Iozzo, R. V. Biologically Active Decorin Is a Monomer in Solution. J. Biol. Chem. 2004, 279, 6606−6612. (270) Santra, M.; Reed, C. C.; Iozzo, R. V. Decorin Binds to a Narrow Region of the Epidermal Growth Factor (Egf) Receptor, Partially Overlapping with but Distinct from the Egf-Binding Epitope. J. Biol. Chem. 2002, 277, 35671−35681. (271) Schlessinger, J. Ligand-Induced, Receptor-Mediated Dimerization and Activation of Egf Receptor. Cell 2002, 110, 669−672. (272) Goldoni, S.; Humphries, A.; Nystrom, A.; Sattar, S.; Owens, R. T.; McQuillan, D. J.; Ireton, K.; Iozzo, R. V. Decorin Is a Novel Antagonistic Ligand of the Met Receptor. J. Cell Biol. 2009, 185, 743− 754. (273) 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. U. S. A. 2013, 110, E2582−E2591. (274) Zhu, J.-X.; Goldoni, S.; Bix, G.; Owens, R. A.; McQuillan, D.; Reed, C. C.; Iozzo, R. V. Decorin Evokes Protracted Internalization and Degradation of the Egf Receptor Via Caveolar Endocytosis. J. Biol. Chem. 2005, 280, 32468−32479. (275) Merline, R.; Nastase, M. V.; Iozzo, R. V.; Schaefer, L. In Extracellular Matrix: Pathobiology and Signaling; Karamanos, N., Ed.; Walter de Gruytier Gmbh and Co.: Berlin, 2012. (276) Schaefer, L.; Iozzo, R. V. Small Leucine-Rich Proteoglycans, at the Crossroad of Cancer Growth and Inflammation. Curr. Opin. Genet. Dev. 2012, 22, 56−57. (277) Morcavallo, A.; Buraschi, S.; Xu, S.-Q.; Belfiore, A.; Schaefer, L.; Iozzo, R. V.; Morrione, A. Decorin Differentially Modulates the

(241) Zoeller, J. J.; Pimtong, W.; Corby, H.; Goldoni, S.; Iozzo, A. E.; Owens, R. T.; Ho, S.-Y.; Iozzo, R. V. A Central Role for Decorin During Vertebrate Convergent Extension. J. Biol. Chem. 2009, 284, 11728−11737. (242) Xu, Y.-Z.; Zhang, Y.-H.; Zhang, Y.-W.; Hong, B.; Liu, J.-M.; Yang, Z.-G. Dynamic Reduction of Plasma Decorin Following Ischemic Stroke: A Pilot Study. Neurochem. Res. 2012, 37, 1843− 1848. (243) Schaefer, L. Small Leucine-Rich Proteoglycans in Kidney Disease. J. Am. Soc. Nephrol. 2011, 22, 1200−1207. (244) Nastase, M. V.; Iozzo, R. V.; Schaefer, L. Key Roles for the Small Leucine-Rich Proteoglycans in Renal and Pulmonary Pathophysiology. Biochim. Biophys. Acta, Gen. Subj. 2014, 1840, 2460−2470. (245) Schaefer, L. Extracellular Matrix Molecules: Endogenous Danger Signals as New Drug Targets in Kidney Diseases. Curr. Opin. Pharmacol. 2010, 10, 185−190. (246) Ichii, M.; Frank, M. B.; Iozzo, R. V.; Kincade, P. W. The Canonical Wnt Pathway Shapes Niches Supportive of Hematopoietic Stem/Progenitor Cells. Blood 2012, 119, 1683−1692. (247) Nandi, A.; Estess, P.; Siegelman, M. H. Hyaluronan Anchoring and Regulation on the Surface of Vascular Endothelial Cells Is Mediated through the Functionally Active Form of Cd44. J. Biol. Chem. 2000, 275, 14939−14948. (248) Robinson, K. A.; Sun, M.; Barnum, C. E.; Weiss, S. N.; Huegel, J.; Shetye, S. S.; Lin, L.; Saez, D.; Adams, S. M.; Iozzo, R. V.; et al. Decorin and Biglycan Are Necessary for Maintaining Collagen Fibril Structure, Fiber Realignment, and Mechanical Properties of Mature Tendons. Matrix Biol. 2017, 64, 81−93. (249) Chen, S.; Birk, D. E. The Regulatory Roles of Small LeucineRich Proteoglycans in Extracellular Matrix Assembly. FEBS J. 2013, 280, 2120−2137. (250) 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-1a and Vascular Endothelial Growth Factor a and Induction of Thrombospondin-1 and Timp3. J. Biol. Chem. 2012, 287, 5492−5506. (251) Neill, T.; Schaefer, L.; Iozzo, R. V. Decorin as a Multivalent Therapeutic Agent against Cancer. Adv. Drug Delivery Rev. 2016, 97, 174−185. (252) Iozzo, R. V.; Schaefer, L. Proteoglycan Form and Function: A Comprehensive Nomenclature of Proteoglycans. Matrix Biol. 2015, 42, 11−55. (253) Patel, A.; Yang, P.; Tinkham, M.; Pradhan, M.; Sun, M. A.; Wang, Y.; Hoang, D.; Wolf, G.; Horton, J. R.; Zhang, X.; et al. DNA Conformation Induces Adaptable Binding by Tandem Zinc Finger Proteins. Cell 2018, 173, 221−233. (254) Vial, C.; Gutierrez, J.; Santander, C.; Cabrera, D.; Brandan, E. Decorin Interacts with Connective Tissue Growth Factor (Ctgf)/ Ccn2 by Lrr12 Inhibiting Its Biological Activity. J. Biol. Chem. 2011, 286, 24242−24252. (255) Khan, G. A.; Girish, G. V.; Lala, N.; DiGuglielmo, G. M.; Lala, P. K. Decorin Is a Novel Vegfr-2-Binding Antagonist for the Human Extravillous Trophoblast. Mol. Endocrinol. 2011, 25, 1431−1443. (256) Kalamajski, S.; Aspberg, A.; Lindblom, K.; Heinegard, D.; Oldberg, A. Asporin Competes with Decorin for Collagen Binding, Binds Calcium and Promotes Osteoblast Collagen Mineralization. Biochem. J. 2009, 423, 53−59. (257) Seidler, D. G. The Galactosaminoglycan-Containing Decorin and Its Impact on Diseases. Curr. Opin. Struct. Biol. 2012, 22, 578− 582. (258) 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, Rev. Cancer 2015, 1855, 276−300. (259) Robinson, P. S.; Lin, T. W.; Reynolds, P. R.; Derwin, K. A.; Iozzo, R. V.; Soslowsky, L. J. Strain-Rate Sensitive Mechanical Properties of Tendon Fascicles from Mice with Genetically BH

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Activity of Insulin Receptor Isoform a Ligands. Matrix Biol. 2014, 35, 82−90. (278) Morrione, A.; Neill, T.; Iozzo, R. V. Dichotomy of Decorin Activity on the Insulin-Like Growth Factor-I System. FEBS J. 2013, 280, 2138−2149. (279) Metalli, D.; Lovat, F.; Tripodi, F.; Genua, M.; Xu, S.-Q.; Spinelli, M.; Alberghina, L.; Vanoni, M.; Baffa, R.; Gomella, L. G.; et al. The Insulin-Like Growth Factor Receptor I Promotes Motility and Invasion of Bladder Cancer Cells through Akt- and MitogenActivated Protein Kinase-Dependent Activation of Paxillin. Am. J. Pathol. 2010, 176, 2997−3006. (280) Skandalis, S. S.; Afratis, N.; Smirlaki, G.; Nikitovic, D.; Theocharis, A. D.; Tzanakakis, G. N.; Karamanos, N. K. Cross-Talk between Estradiol Receptor and Egfr/Igf-Ir Signaling Pathways in Estrogen-Responsive Breast Cancers: Focus on the Role and Impact of Proteoglycans. Matrix Biol. 2014, 35, 182−193. (281) Piperigkou, Z.; Franchi, M.; Gotte, M.; Karamanos, N. K. Estrogen Receptor Beta as Epigenetic Mediator of Mir-10b and Mir145 in Mammary Cancer. Matrix Biol. 2017, 64, 94−111. (282) Schlessinger, J. Common and Distinct Elements in Cellular Signaling Via Egf and Fgf Receptors. Science 2004, 306, 1506−1507. (283) Lemmon, M. A.; Schlessinger, J. Cell Signaling by Receptor Tyrosine Kinases. Cell 2010, 141, 1117−1134. (284) Moscatello, D. K.; Santra, M.; Mann, D. M.; McQuillan, D. J.; Wong, A. J.; Iozzo, R. V. Decorin Suppresses Tumor Cell Growth by Activating the Epidermal Growth Factor Receptor. J. Clin. Invest. 1998, 101, 406−412. (285) Csordas, G.; Santra, M.; Reed, C. C.; Eichstetter, I.; McQuillan, D. J.; Gross, D.; Nugent, M. A.; Hajnoczky, G.; Iozzo, R. V. Sustained Down-Regulation of the Epidermal Growth Factor Receptor by Decorin. A Mechanism for Controlling Tumor Growth in Vivo. J. Biol. Chem. 2000, 275, 32879−32887. (286) Santra, M.; Eichstetter, I.; Iozzo, R. V. An Anti-Oncogenic Role for Decorin: Downregulation of Erbb2 Leads to Growth Suppression and Cytodifferentiation of Mammary Carcinoma Cells. J. Biol. Chem. 2000, 275, 35153−35161. (287) Minor, K. H.; Bournat, J. C.; Toscano, N.; Giger, R. J.; Davies, S. J. A. Decorin, Erythroblastic Leukaemia Viral Oncogene Homologue B4 and Signal Transducer and Activator of Transcription 3 Regulation of Semaphorin 3a in Central Nervous System Scar Tissue. Brain 2011, 134, 1140−1155. (288) Seidler, D. G.; Goldoni, S.; Agnew, C.; Cardi, C.; Thakur, M. L.; Owens, R. A.; McQuillan, D. J.; Iozzo, R. V. Decorin Protein Core Inhibits in Vivo Cancer Growth and Metabolism by Hindering Epidermal Growth Factor Receptor Function and Triggering Apoptosis Via Caspase-3 Activation. J. Biol. Chem. 2006, 281, 26408−26418. (289) Abulrob, A.; Giuseppin, S.; Andrade, M. F.; McDermid, A.; Moreno, M.; Stanimirovic, D. Interactions of Egfr and Caveolin-1 in Human Glioblastoma Cells: Evidence That Tyrosine Phosphorylation Regulates Egfr Association with Caveolae. Oncogene 2004, 23, 6967− 6979. (290) Buraschi, S.; Pal, N.; Tyler-Rubinstein, N.; Owens, R. T.; Neill, T.; Iozzo, R. V. Decorin Antagonizes Met Receptor Activity and Downregulates B-Catenin and Myc Levels. J. Biol. Chem. 2010, 285, 42075−42085. (291) Aberle, H.; Bauer, A.; Stappert, J.; Kispert, A.; Kemler, R. BCatenin Is a Target for the Ubiquitin-Proteasome Pathway. EMBO J. 1997, 16, 3797−3804. (292) Danilkovitch-Miagkova, A.; Miagkov, A.; Skeel, A.; Nakaigawa, N.; Zbar, B.; Leonard, E. J. Oncogenic Mutants of Ron and Met Receptor Tyrosine Kinases Cause Activation of the BCatenin Pathway. Mol. Cell. Biol. 2001, 21, 5857−5868. (293) Clevers, H. Wnt/B-Catenin Signaling in Development and Disease. Cell 2006, 127, 469−480. (294) Monga, S. P. S.; Mars, W. M.; Pediaditakis, P.; Bell, A.; Mule, K.; Bowen, W. C.; Wang, X.; Zarnegar, R.; Michalopoulos, G. K. Hepatocyte Growth Factor Induces Wnt-Independent Nuclear

Translocation of B-Catenin after Met-B-Catenin Dissociation in Hepatocytes. Cancer Res. 2002, 62, 2064−2071. (295) Rasola, A.; Fassetta, M.; De Bacco, F.; D’Alessandro, L.; Gramaglia, D.; Di Renzo, M. G.; Comoglio, P. M. A Positive Feedback Loop between Hepatocyte Growth Factor Receptor and BCatenin Sustains Colorectal Cancer Cell Invasive Growth. Oncogene 2007, 26, 1078−1087. (296) Finch, A. J.; Soucek, L.; Junttila, M. R.; Swigart, L. B.; Evan, G. I. Acute Overexpression of Myc in Intestinal Epithelium Recapitulates Some but Not All the Changes Elicited by Wnt/B-Catenin Pathway Activation. Mol. Cell. Biol. 2009, 29, 5306−5315. (297) Herynk, M. H.; Tsan, R.; Radinsky, R.; Gallick, G. E. Activation of C-Met in Colorectal Carcinoma Cells Leads to Constitutive Association of Tyrosine-Phosphorylated B-Catenin. Clin. Exp. Metastasis 2003, 20, 291−300. (298) Kessenbrock, K.; Wang, C.-Y.; Werb, Z. Matrix Metalloproteinases in Stem Cell Mobilization. Matrix Biol. 2015, 44−46, 184−190. (299) Wells, J. M.; Gaggar, A.; Blalock, J. E. Mmp Generated Matrikines. Matrix Biol. 2015, 44−46, 122−129. (300) Arpino, V.; Brock, M.; Gill, S. E. The Role of Timps in Regulation of Extracellular Matrix Proteolysis. Matrix Biol. 2015, 44− 46, 247−254. (301) Vadon-Le Goff, S.; Hulmes, D. J.; Moali, C. Bmp-1/TolloidLike Proteinases Synchronize Matrix Assembly with Growth Factor Activation to Promote Morphogenesis and Tissue Remodeling. Matrix Biol. 2015, 44−46, 14−23. (302) Deryugina, E. I.; Quigley, J. P. Tumor Angiogenesis: MmpMediated Induction of Intravasation- and Metastasis-Sustaining Neovasculature. Matrix Biol. 2015, 44−46, 94−112. (303) Apte, S. S.; Parks, W. C. Metalloproteinases: A Parade of Functions in Matrix Biology and an Outlook for the Future. Matrix Biol. 2015, 44−46, 1−6. (304) Rohani, M. G.; Parks, W. C. Matrix Remodelling by Mmps During Wound Repair. Matrix Biol. 2015, 44−46, 113−121. (305) Itoh, Y. Membrane-Type Matrix Metalloproteinases: Their Functions and Regulations. Matrix Biol. 2015, 44−46, 207−223. (306) Hubmacher, D.; Apte, S. S. Adamts Proteins as Modulators of Microfibril Formation and Function. Matrix Biol. 2015, 47, 34−43. (307) Scilabra, S. D.; Yamamoto, K.; Pigoni, M.; Sakamoto, K.; Muller, S. A.; Papadopoulou, A.; Lichtenthaler, S. F.; Troeberg, L.; Nagase, H.; Kadomatsu, K. Dissecting the Interaction between Tissue Inhibitor of Metalloproteinases-3 (Timp-3) and Low Density Lipoprotein Receptor-Related Protein-1 (Lrp-1): Development of a ″Trap″ to Increase Levels of Timp-3 in the Tissue. Matrix Biol. 2017, 59, 69−79. (308) Andreuzzi, E.; Colladel, R.; Pellicani, R.; Tarticchio, G.; Cannizzaro, R.; Spessotto, P.; Bussolati, B.; Brossa, A.; De Paoli, P.; Canzonieri, V.; et al. The Angiostatic Molecule Multimerin 2 Is Processed by Mmp-9 to Allow Sprouting Angiogenesis. Matrix Biol. 2017, 64, 40−53. (309) Galvagni, F.; Nardi, F.; Spiga, O.; Trezza, A.; Tarticchio, G.; Pellicani, R.; Andreuzzi, E.; Caldi, E.; Toti, P.; Tosi, G. M.; et al. Dissecting the Cd93-Multimerin 2 Interaction Involved in Cell Adhesion and Migration of the Activated Endothelium. Matrix Biol. 2017, 64, 112−127. (310) Murphy-Ullrich, J. E.; Sage, E. H. Revisiting the Matricellular Concept. Matrix Biol. 2014, 37, 1−14. (311) Wallace, D. M.; Murphy-Ullrich, J. E.; Downs, J. C.; O’Brien, C. J. The Role of Matricellular Proteins in Glaucoma. Matrix Biol. 2014, 37, 174−182. (312) Resovi, A.; Pinessi, D.; Chiorino, G.; Taraboletti, G. Current Understanding of the Thrombospondin-1 Interactome. Matrix Biol. 2014, 37, 83−91. (313) Rogers, N. M.; Sharifi-Sanjani, M.; Csanyi, G.; Pagano, P. J.; Isenberg, J. S. Thrombospondin-1 and Cd47 Regulation of Cardiac, Pulmonary and Vascular Responses in Health and Disease. Matrix Biol. 2014, 37, 92−101. BI

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(314) Duquette, M.; Nadler, M.; Okuhara, D.; Thompson, J.; Shuttleworth, T.; Lawler, J. Members of the Thrombospondin Gene Family Bind Stromal Interaction Molecule 1 and Regulate Calcium Channel Activity. Matrix Biol. 2014, 37, 15−24. (315) Soto-Pantoja, D. R.; Shih, H. B.; Maxhimer, J. B.; Cook, K. L.; Ghosh, A.; Isenberg, J. S.; Roberts, D. D. Thrombospondin-1 and Cd47 Signaling Regulate Healing of Thermal Injury in Mice. Matrix Biol. 2014, 37, 25−34. (316) MacLauchlan, S. C.; Calabro, N. E.; Huang, Y.; Krishna, M.; Bancroft, T.; Sharma, T.; Yu, J.; Sessa, W. C.; Giordano, F.; Kyriakides, T. R. Hif-1a Represses the Expression of the Angiogenesis Inhibitor Thrombospondin-2. Matrix Biol. 2018, 65, 45−58. (317) Buraschi, S.; Neill, T.; Owens, R. T.; Iniguez, L. A.; Purkins, G.; Vadigepalli, R.; Evans, B.; Schaefer, L.; Peiper, S. C.; Wang, Z.; et al. Decorin Protein Core Affects the Global Gene Expression Profile of the Tumor Microenvironment in a Triple-Negative Orthotopic Breast Carcinoma Xenograft Model. PLoS One 2012, 7, e45559. (318) Relaix, F.; Weng, X.; Marazzi, G.; Yang, E.; Copeland, N.; Jenkins, N.; Spence, S. E.; Sassoon, D. Pw1, a Novel Zinc Finger Gene Implicated in the Myogenic and Neuronal Lineages. Dev. Biol. 1996, 177, 383−396. (319) Kim, J.; Ashworth, L.; Branscomb, E.; Stubbs, L. The Human Homolog of a Mouse-Imprinted Gene, Peg3, Maps to a Zinc Finger Gene-Rich Region of Human Chromosome 19q13.4. Genome Res. 1997, 7, 532−540. (320) Kuroiwa, Y.; Kaneko-Ishino, T.; Kagitani, F.; Kohda, T.; Li, L.L.; Tada, M.; Suzuki, R.; Yokoyama, M.; Shiroishi, T.; Wakana, S.; et al. Peg3 Imprinted Gene on Proximal Chromosome 7 Encodes for a Zinc Finger Protein. Nat. Genet. 1996, 12, 186−190. (321) Thiaville, M. M.; Huang, J. M.; Kim, H.; Ekram, M. B.; Roh, T.-Y.; Kim, J. DNA-Binding Motif and Target Genes of the Imprinted Transcription Factor Peg3. Gene 2013, 512, 314−320. (322) Besson, V.; Smeriglio, P.; Wegener, A.; Relaix, F.; Nait Oumesmar, B.; Sassoon, D. A.; Marazzi, G. Pw1 Gene/Paternally Expressed Gene 3 (Pw1/Peg3) Identifies Multiple Adult Stem and Progenitor Cell Populations. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 11470−11475. (323) Bonfanti, C.; Rossi, G.; Tedesco, F. S.; Giannotta, M.; Benedetti, S.; Tonlorenzi, R.; Antonini, S.; Marazzi, G.; Dejana, E.; Sassoon, D.; et al. Pw1/Peg3 Expression Regulates Key Properties That Determine Mesoangioblast Stem Cell Competence. Nat. Commun. 2015, 6, 6364. (324) Relaix, F.; Wei, X.-j.; Li, W.; Pan, J.; Lin, Y.; Bowtell, D. D.; Sassoon, D. A.; Wu, X. Pw1/Peg3 Is a Potential Cell Death Mediator and Cooperates with Siah1a in P53-Mediated Apoptosis. Proc. Natl. Acad. Sci. U. S. A. 2000, 97, 2105−2110. (325) Johnson, M. D.; Wu, X.; Aithmitti, N.; Morrison, R. S. Peg3/ Pw1 Is a Mediator between P53 and Bax in DNA Damage-Induced Neuronal Death. J. Biol. Chem. 2002, 277, 23000−23007. (326) Yamaguchi, A.; Taniguchi, M.; Hori, O.; Ogawa, S.; Tojo, N.; Matsuoka, N.; Miyake, S.-i.; Kasai, K.; Sugimoto, H.; Tamatani, M.; et al. Peg3/Pw1 Is Involved in P53-Mediated Cell Death Pathway in Brain Ischemia/Hypoxia. J. Biol. Chem. 2002, 277, 623−629. (327) Deng, Y.; Wu, X. Peg3/Pw1 Promotes P53-Mediated Apoptosis by Inducing Bax Translocation from Cytosol to Mitochondria. Proc. Natl. Acad. Sci. U. S. A. 2000, 97, 12050−12055. (328) Coletti, D.; Yang, E.; Marazzi, G.; Sassoon, D. Tnf alpha Inhibits Skeletal Myogenesis through a Pw1-Dependent Pathway by Recruitment of Caspase Pathways. EMBO J. 2002, 21, 631−642. (329) Li, L.-L.; Keverne, E. B.; Aparicio, S. A.; Ishino, F.; Barton, S. C.; Surani, M. A. Regulation of Maternal Behavior and Offspring Growth by Paternally Expressed Peg3. Science 1999, 284, 330−334. (330) Champagne, F. A.; Curley, J. P.; Swaney, W. T.; Hasen, N. S.; Keverne, E. B. Paternal Influence on Female Behavior: The Role of Peg3 in Exploration, Olfaction, and Neuroendocrine Regulation of Maternal Behavior of Female Mice. Behav. Neurosci. 2009, 123, 469− 480.

(331) Kohda, T.; Asai, A.; Kuroiwa, Y.; Kobayashi, S.; Aisaka, K.; Nagashima, G.; Yoshida, M. C.; Kondo, Y.; Kagiyama, N.; Kirino, T.; et al. Tumour Suppressor Activity of Human Imprinted Gene Peg3 in a Glioma Cell Line. Genes Cells 2001, 6, 237−247. (332) Nye, M. D.; Hoyo, C.; Huang, Z.; Vidal, A. C.; Wang, F.; Overcash, F.; Smith, J. S.; Vasquez, B.; Hernandez, B.; Swai, B.; Oneko, O.; et al. Association between Methylation of Paternally Expressed Gene 3 (Peg3), Cervical Intraepithelial Neoplasia and Invasive Cervical Cancer. PLoS One 2013, 8, e56325. (333) Dowdy, S. C.; Gostout, B. S.; Shridhar, V.; Wu, X.; Smith, D. I.; Podratz, K. C.; Jiang, S.-W. Biallelic Methylation and Silencing of Paternally Expressed Gene 3 (Peg3) in Gynecologic Cancer Cell Lines. Gynecol. Oncol. 2005, 99, 126−134. (334) Feng, W.; Marquez, R. T.; Lu, Z.; Liu, J.; Lu, K. H.; Issa, J.-P. J.; Fishman, D. M.; Yu, Y.; Bast, R. C. Imprinted Tumor Suppressor Genes Arhi and Peg3 Are the Most Frequently Down-Regulated in Human Ovarian Cancers by Loss of Heterozygosity and Promoter Methylation. Cancer 2008, 112, 1489−1502. (335) Maegawa, S.; Yoshioka, H.; Itaba, N.; Kubota, N.; Nishihara, S.; Shirayoshi, Y.; Nanba, E.; Oshimura, M. Epigenetic Silencing of Peg3 Gene Expression in Human Glioma Cell Lines. Mol. Carcinog. 2001, 31, 1−9. (336) Jiang, X.; Yu, Y.; Yang, H. W.; Agar, N. Y. R.; Frado, L.; Johnson, M. D. The Imprinted Gene Peg3 Inhibits Wnt Signaling and Regulates Glioma Growth. J. Biol. Chem. 2010, 285, 8472−8480. (337) Mizushima, N.; Levine, B. Autophagy in Mammalian Development and Differentiation. Nat. Cell Biol. 2010, 12, 823−830. (338) He, C.; Klionsky, D. J. Regulation Mechanisms and Signaling Pathways of Autophagy. Annu. Rev. Genet. 2009, 43, 67−93. (339) Neill, T.; Torres, A. T.; Buraschi, S.; Iozzo, R. V. Decorin Has an Appetite for Endothelial Cell Autophagy. Autophagy 2013, 9, 1626−1628. (340) Torres, A.; Gubbiotti, M. A.; Iozzo, R. V. Decorin-Inducible Peg3 Evokes Beclin 1-Mediated Autophagy and Thrombospondin 1Mediated Angiostasis. J. Biol. Chem. 2017, 292, 5055−5069. (341) Neill, T.; Sharpe, C.; Owens, R. T.; Iozzo, R. V. DecorinEvoked Paternally Expressed Gene 3 (Peg3) Is an Upstream Regulator of the Transcription Factor Eb (Tfeb) in Endothelial Cell Autophagy. J. Biol. Chem. 2017, 292, 16211−16220. (342) Pattingre, S.; Levine, B. Bcl-2 Inhibition of Autophagy: A New Route to Cancer? Cancer Res. 2006, 66, 2885−2888. (343) Fong, T. A. T.; Shawver, L. K.; Sun, L.; Tang, C.; App, H.; Powell, T. J.; Kim, Y. H.; Schreck, R.; Wang, X.; Risau, W.; et al. Su5416 Is a Potent and Selective Inhibitor of the Vascular Endothelial Growth Factor Receptor (Flk-1/Kdr) That Inhibits Tyrosine Kinase Catalysis, Tumor Vascularization, and Growth of Multiple Tumor Types. Cancer Res. 1999, 59, 99−106. (344) 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, 34, 46−54. (345) 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. (346) Fullgrabe, J.; Klionsky, D. J.; Joseph, B. The Return of the Nucleus: Transcriptional and Epigenetic Control of Autophagy. Nat. Rev. Mol. Cell Biol. 2014, 15, 65−74. (347) Settembre, C.; Ballabio, A. Tfeb Regulates Autophagy: An Integrated Coordination of Cellular Degradation and Recycling Processes. Autophagy 2011, 7, 1379−1381. (348) Settembre, C.; Fraldi, A.; Medina, D. L.; Ballabio, A. Signals from the Lysosome: A Control Centre for Cellular Clearance and Energy Metabolism. Nat. Rev. Mol. Cell Biol. 2013, 14, 283−296. (349) Settembre, C.; Di Malta, C.; Polito, V. A.; Arencibia, M. G.; Vetrini, F.; Erdin, S.; Huynh, T.; Medina, D.; Colella, P.; Sardiello, M.; et al. Tfeb Links Autophagy to Lysosomal Biogenesis. Science 2011, 332, 1429−1433. BJ

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Turnover in Vivo. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, 6400− 6405. (367) Aita, V. M.; Liang, X. H.; Murty, V. V.; Pincus, D. L.; Yu, W.; Cayanis, E.; Kalachikov, S.; Gilliam, T. C.; Levine, B. Cloning and Genomic Organization of Beclin 1, a Candidate Tumor Suppressor Gene on Chromosome 17q21. Genomics 1999, 59, 59−65. (368) Matsunaga, K.; Saitoh, T.; Tabata, K.; Omori, H.; Satoh, T.; Kurotori, N.; Maejima, I.; Shirahama-Noda, K.; Ichimura, T.; Isobe, T.; et al. Two Beclin 1-Binding Proteins, Atg14l and Rubicon, Reciprocally Regulate Autophagy at Different Stages. Nat. Cell Biol. 2009, 11, 385−396. (369) Yue, Z.; Jin, S.; Yang, C.; Levine, A. J.; Heintz, N. Beclin 1, an Authophagy Gene Essential for Early Embryonic Development, Is a Haploinsufficient Tumor Suppressor. Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 15077−15082. (370) Stjepanovic, G.; Baskaran, S.; Lin, M. G.; Hurley, J. H. Vps34 Kinase Domain Dynamics Regulate the Autophagic Pi 3-Kinase Complex. Mol. Cell 2017, 67, 528−534. (371) Funderburk, S. F.; Wang, Q. J.; Yue, Z. The Beclin 1-Vps34 Complex- at the Crossroads of Autophagy and Beyond. Trends Cell Biol. 2010, 20, 355−362. (372) Kane, L. A.; Lazarou, M.; Fogel, A. I.; Li, Y.; Yamano, K.; Sarraf, S. A.; Banerjee, S.; Youle, R. J. Pink1 Phosphorylates Ubiquitin to Activate Parkin E3 Ubiquitin Ligase Activity. J. Cell Biol. 2014, 205, 143−153. (373) Koyano, F.; Okatsu, K.; Kosako, H.; Tamura, Y.; Go, E.; Kimura, M.; Kimura, Y.; Tsuchiya, H.; Yoshihara, H.; Hirokawa, T.; et al. Ubiquitin Is Phosphorylated by Pink1 to Activate Parkin. Nature 2014, 510, 162−166. (374) Matsuda, N.; Sato, S.; Shiba, K.; Okatsu, K.; Saisho, K.; Gautier, C. A.; Sou, Y. S.; Saiki, S.; Kawajiri, S.; Sato, F.; et al. Pink1 Stabilized by Mitochondrial Depolarization Recruits Parkin to Damaged Mitochondria and Activates Latent Parkin for Mitophagy. J. Cell Biol. 2010, 189, 211−221. (375) Iguchi, M.; Kujuro, Y.; Okatsu, K.; Koyano, F.; Kosako, H.; Kimura, M.; Suzuki, N.; Uchiyama, S.; Tanaka, K.; Matsuda, N. Parkin-Catalyzed Ubiquitin-Ester Transfer Is Triggered by Pink1Dependent Phosphorylation. J. Biol. Chem. 2013, 288, 22019−22032. (376) Geisler, S.; Holmstrom, K. M.; Skujat, D.; Fiesel, F. C.; Rothfuss, O. C.; Kahle, P. J.; Springer, W. Pink1/Parkin-Mediated Mitophagy Is Dependent on Vdac1 and P62/Sqstm1. Nat. Cell Biol. 2010, 12, 119−131. (377) Xu, W.; Neill, T.; Yang, Y.; Hu, Z.; Cleveland, E.; Wu, Y.; Hutten, R.; Xiao, X.; Stock, S. R.; Shevrin, D.; et al. The Systemic Delivery of an Oncolytic Adenovirus Expressing Decorin Inhibits Bone Metastasis in a Mouse Model of Human Prostate Cancer. Gene Ther. 2015, 22, 247−256. (378) Hassell, J. R.; Newsome, D. A.; Krachmer, J. H.; Rodrigues, M. M. Macular Corneal Dystrophy: Failure to Synthesize a Mature Keratan Sulfate Proteoglycan. Proc. Natl. Acad. Sci. U. S. A. 1980, 77, 3705−3709. (379) Nikitovic, D.; Katonis, P.; Tsatsakis, A.; Karamanos, N. K.; Tzanakakis, G. N. Lumican, a Small Leucine-Rich Proteoglycan. IUBMB Life 2008, 60, 818−823. (380) Iozzo, R. V. The Family of the Small Leucine-Rich Proteoglycans: Key Regulators of Matrix Assembly and Cellular Growth. Crit. Rev. Biochem. Mol. Biol. 1997, 32, 141−174. (381) Grover, J.; Chen, X. N.; Korenberg, J. R.; Roughley, P. J. The Human Lumican Gene. Organization, Chromosomal Location, and Expression in Articular Cartilage. J. Biol. Chem. 1995, 270, 21942− 21949. (382) Corpuz, L. M.; Funderburgh, J. L.; Funderburgh, M. L.; Bottomley, G. S.; Prakash, S.; Conrad, G. W. Molecular Cloning and Tissue Distribution of Keratocan. Bovine Corneal Keratan Sulfate Proteoglycan 37a. J. Biol. Chem. 1996, 271, 9759−9763. (383) Cornuet, P. K.; Blochberger, T. C.; Hassell, J. R. Molecular Polymorphism of Lumican During Corneal Development. Invest. Ophthalmol. Vis. Sci. 1994, 35, 870−877.

(350) Palmieri, M.; Impey, S.; Kang, H.; di Ronza, A.; Pelz, C.; Sardiello, M.; Ballabio, A. Characterization of the Clear Network Reveals an Integrated Control of Cellular Clearance Pathways. Hum. Mol. Genet. 2011, 20, 3852−3866. (351) Leeman, D. S.; Hebestreit, K.; Ruetz, T.; Webb, A. E.; McKay, A.; Pollina, E. A.; Dulken, B. W.; Zhao, X.; Yeo, R. W.; Ho, T. T.; et al. Lysosome Activation Clears Aggregates and Enhances Quiescent Neural Stem Cell Activation During Aging. Science 2018, 359, 1277− 1283. (352) Gubbiotti, M. A.; Neill, T.; Frey, H.; Schaefer, L.; Iozzo, R. V. Decorin Is an Autophagy-Inducible Proteoglycan and Is Required for Proper in Vivo Autophagy. Matrix Biol. 2015, 48, 14−25. (353) Gubbiotti, M. A.; Iozzo, R. V. Proteoglycans Regulate Autophagy Via Outside-in Signaling: An Emerging New Concept. Matrix Biol. 2015, 48, 6−13. (354) Wei, Y.; Zou, Z.; Becker, N.; Anderson, M.; Sumpter, R.; Xiao, G.; Kinch, L.; Koduru, P.; Christudass, C. S.; Veltri, R. W.; et al. EgfrMediated Beclin 1 Phosphorylation in Autophagy Suppression, Tumor Progression, and Tumor Chemoresistance. Cell 2013, 154, 1269−1284. (355) Wang, R. C.; Wei, Y.; An, Z.; Zou, Z.; Xiao, G.; Bhagat, G.; White, M.; Reichelt, J.; Levine, B. Akt-Mediated Regulation of Autophagy and Tumorigenesis through Beclin 1 Phosphorylation. Science 2012, 338, 956−959. (356) Neill, T.; Torres, A.; Buraschi, S.; Owens, R. T.; Hoek, J.; Baffa, R.; Iozzo, R. V. Decorin Induces Mitophagy in Breast Carcinoma Cells Via Peroxisome Proliferator-Activated Receptor G Coactivator-1a (Pgc-1a) and Mitostatin. J. Biol. Chem. 2014, 289, 4952−4968. (357) Vecchione, A.; Fassan, M.; Anesti, V.; Morrione, A.; Goldoni, S.; Baldassarre, G.; Byrne, D.; D’Arca, D.; Palazzo, J. P.; Lloyd, J.; et al. Mitostatin, a Putative Tumor Suppressor on Chromosome 12q24.1, Is Downregulated in Human Bladder and Breast Cancer. Oncogene 2009, 28, 257−269. (358) Fassan, M.; D’Arca, D.; Letko, J.; Vecchione, A.; Gardiman, M. P.; McCue, P.; Wildemore, B.; Rugge, M.; Shupp-Byrne, D.; Gomella, L. G.; et al. Mitostatin Is Down-Regulated in Human Prostate Cancer and Suppresses the Invasive Phenotype of Prostate Cancer Cells. PLoS One 2011, 6, e19771. (359) Cerqua, C.; Anesti, V.; Pyakurel, A.; Liu, D.; Naon, D.; Wiche, G.; Baffa, R.; Dimmer, K. S.; Scorrano, L. Trichoplein/Mitostatin Regulates Endoplasmic Reticulum-Mitochondria Juxtaposition. EMBO Rep. 2010, 11, 854−860. (360) Ventura-Clapier, R.; Garnier, A.; Veksler, W. Transcriptional Control of Mitochondrial Biogenesis: The Central Role of Pgc-1a. Cardiovasc. Res. 2008, 79, 208−217. (361) Haq, R.; Shoag, J.; Andreu-Perez, P.; Yokoyama, S.; Edelman, H.; Rowe, G. C.; Frederick, D. T.; Hurley, A. D.; Nellore, A.; Kung, A. L.; et al. Oncogenic Braf Regulates Oxidative Metabolism Via Pgc1a and Mitf. Cancer Cell 2013, 23, 302−315. (362) Vazquez, F.; Lim, J.-H.; Chim, H.; Bhalla, K.; Girnun, G.; Pierce, K.; Clish, C. B.; Granter, S. R.; Widlund, H. R.; Spiegelman, B. M.; et al. Pgc1a Expression Defines a Subset of Human Melanoma Tumors with Increased Mitochondrial Capacity and Resistance to Oxidative Stress. Cancer Cell 2013, 23, 287−301. (363) Saito, T.; Sadoshima, J. Molecular Mechanisms of Mitochondrial Autophagy/Mitophagy in the Heart. Circ. Res. 2015, 116, 1477−1490. (364) Dagda, R.; Cherra, S. J. I.; Kulich, S. M.; Tandon, A.; Park, D.; Chu, C. T. Loss of Pink1 Function Promotes Mitophagy through Effects on Oxidative Stress and Mitochondrial Fission. J. Biol. Chem. 2009, 284, 13843−13855. (365) Kubli, D. A.; Gustafsson, A. B. Mitochondria and Mitophagy: The Yin and Yang of Cell Death Control. Circ. Res. 2012, 111, 1208− 1221. (366) Vincow, E. S.; Merrihew, G.; Thomas, R. E.; Shulman, N. J.; Beyer, R. P.; MacCoss, M. J.; Pallanck, L. J. The Pink1-Parkin Pathway Promotes Both Mitophagy and Selective Respiratory Chain BK

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Adenocarcinoma and Squamous Cell Carcinoma. Int. J. Oncol. 2008, 33, 1177−1185. (400) Yamamoto, T.; Matsuda, Y.; Kawahara, K.; Ishiwata, T.; Naito, Z. Secreted 70kda Lumican Stimulates Growth and Inhibits Invasion of Human Pancreatic Cancer. Cancer Lett. 2012, 320, 31−39. (401) Lu, Y. P.; Ishiwata, T.; Kawahara, K.; Watanabe, M.; Naito, Z.; Moriyama, Y.; Sugisaki, Y.; Asano, G. Expression of Lumican in Human Colorectal Cancer Cells. Pathol. Int. 2002, 52, 519−526. (402) Radwanska, A.; Litwin, M.; Nowak, D.; Baczynska, D.; Wegrowski, Y.; Maquart, F. X.; Malicka-Blaszkiewicz, M. Overexpression of Lumican Affects the Migration of Human Colon Cancer Cells through up-Regulation of Gelsolin and Filamentous Actin Reorganization. Exp. Cell Res. 2012, 318, 2312−2323. (403) Nikitovic, D.; Chatzinikolaou, G.; Tsiaoussis, J.; Tsatsakis, A.; Karamanos, N. K.; Tzanakakis, G. N. Insights into Targeting Colon Cancer Cell Fate at the Level of Proteoglycans/Glycosaminoglycans. Curr. Med. Chem. 2012, 19, 4247−4258. (404) Nikitovic, D.; Chalkiadaki, G.; Berdiaki, A.; Aggelidakis, J.; Katonis, P.; Karamanos, N. K.; Tzanakakis, G. N. Lumican Regulates Osteosarcoma Cell Adhesion by Modulating Tgfbeta2 Activity. Int. J. Biochem. Cell Biol. 2011, 43, 928−935. (405) Pietraszek, K.; Brezillon, S.; Perreau, C.; Malicka-Blaszkiewicz, M.; Maquart, F. X.; Wegrowski, Y. Lumican - Derived Peptides Inhibit Melanoma Cell Growth and Migration. PLoS One 2013, 8, e76232. (406) Pietraszek, K.; Chatron-Colliet, A.; Brezillon, S.; Perreau, C.; Jakubiak-Augustyn, A.; Krotkiewski, H.; Maquart, F. X.; Wegrowski, Y. Lumican: A New Inhibitor of Matrix Metalloproteinase-14 Activity. FEBS Lett. 2014, 588, 4319−4324. (407) Stasiak, M.; Boncela, J.; Perreau, C.; Karamanou, K.; ChatronColliet, A.; Proult, I.; Przygodzka, P.; Chakravarti, S.; Maquart, F. X.; Kowalska, M. A.; et al. Lumican Inhibits Snail-Induced Melanoma Cell Migration Specifically by Blocking Mmp-14 Activity. PLoS One 2016, 11, e0150226. (408) Pietraszek-Gremplewicz, K.; Karamanou, K.; Niang, A.; Dauchez, M.; Belloy, N.; Maquart, F. X.; Baud, S.; Brezillon, S. Small Leucine-Rich Proteoglycans and Matrix Metalloproteinase-14: Key Partners? Matrix Biol. 2017, DOI: 10.1016/j.matbio.2017.12.006. (409) Chakravarti, S.; Stallings, R. L.; SundarRaj, N.; Cornuet, P. K.; Hassell, J. R. Primary Structure of Human Lumican (Keratan Sulfate Proteoglycan) and Localization of the Gene (Lum) to Chromosome 12q21.3-Q22. Genomics 1995, 27, 481−488. (410) Funderburgh, J. L.; Funderburgh, M. L.; Brown, S. J.; Vergnes, J. P.; Hassell, J. R.; Mann, M. M.; Conrad, G. W. Sequence and Structural Implications of a Bovine Corneal Keratan Sulfate Proteoglycan Core Protein. Protein 37b Represents Bovine Lumican and Proteins 37a and 25 Are Unique. J. Biol. Chem. 1993, 268, 11874−11880. (411) Blochberger, T. C.; Cornuet, P. K.; Hassell, J. R. Isolation and Partial Characterization of Lumican and Decorin from Adult Chicken Corneas. A Keratan Sulfate-Containing Isoform of Decorin Is Developmentally Regulated. J. Biol. Chem. 1992, 267, 20613−20619. (412) Funderburgh, J. L.; Funderburgh, M. L.; Hevelone, N. D.; Stech, M. E.; Justice, M. J.; Liu, C. Y.; Kao, W. W.; Conrad, G. W. Sequence, Molecular Properties, and Chromosomal Mapping of Mouse Lumican. Invest. Ophthalmol. Vis. Sci. 1995, 36, 2296−2303. (413) Scott, P. G.; McEwan, P. A.; Dodd, C. M.; Bergmann, E. M.; Bishop, P. N.; Bella, J. Crystal Structure of the Dimeric Protein Core of Decorin, the Archetypal Small Leucine-Rich Repeat Proteoglycan. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 15633−15638. (414) Scott, P. G.; Dodd, C. M.; Bergmann, E. M.; Sheehan, J. K.; Bishop, P. N. Crystal Structure of the Biglycan Dimer and Evidence That Dimerization Is Essential for Folding and Stability of Class I Small Leucine-Rich Repeat Proteoglycans. J. Biol. Chem. 2006, 281, 13324−13332. (415) Yamanaka, O.; Yuan, Y.; Coulson-Thomas, V. J.; Gesteira, T. F.; Call, M. K.; Zhang, Y.; Zhang, J.; Chang, S. H.; Xie, C.; Liu, C. Y.; et al. Lumican Binds Alk5 to Promote Epithelium Wound Healing. PLoS One 2013, 8, e82730.

(384) Brezillon, S.; Pietraszek, K.; Maquart, F. X.; Wegrowski, Y. Lumican Effects in the Control of Tumour Progression and Their Links with Metalloproteinases and Integrins. FEBS J. 2013, 280, 2369−2381. (385) Ping Lu, Y.; Ishiwata, T.; Asano, G. Lumican Expression in Alpha Cells of Islets in Pancreas and Pancreatic Cancer Cells. J. Pathol. 2002, 196, 324−330. (386) Leygue, E.; Snell, L.; Dotzlaw, H.; Hole, K.; Hiller-Hitchcock, T.; Roughley, P. J.; Watson, P. H.; Murphy, L. C. Expression of Lumican in Human Breast Carcinoma. Cancer Res. 1998, 58, 1348− 1352. (387) Leygue, E.; Snell, L.; Dotzlaw, H.; Troup, S.; Hiller-Hitchcock, T.; Murphy, L. C.; Roughley, P. J.; Watson, P. H. Lumican and Decorin Are Differentially Expressed in Human Breast Carcinoma. J. Pathol. 2000, 192, 313−320. (388) Troup, S.; Njue, C.; Kliewer, E. V.; Parisien, M.; Roskelley, C.; Chakravarti, S.; Roughley, P. J.; Murphy, L. C.; Watson, P. H. Reduced Expression of the Small Leucine-Rich Proteoglycans, Lumican, and Decorin Is Associated with Poor Outcome in NodeNegative Invasive Breast Cancer. Clin. Cancer Res. 2003, 9, 207−214. (389) Vuillermoz, B.; Khoruzhenko, A.; D’Onofrio, M.-F.; Ramont, L.; Venteo, L.; Perreau, C.; Antonicelli, F.; Maquart, F.-X.; Wegrowski, Y. The Small Leucine-Rich Proteoglycan Lumican Inhibits Melanoma Progression. Exp. Cell Res. 2004, 296, 294−306. (390) Brezillon, S.; Zeltz, C.; Schneider, L.; Terryn, C.; Vuillermoz, B.; Ramont, L.; Perrau, C.; Pluot, M.; Diebold, M. D.; Radwanska, A.; et al. Lumican Inhibits B16f1 Melanoma Cell Lung Metastasis. J. Physiol. Pharmacol. 2009, 60, 15−22. (391) D’Onofrio, M. F.; Brezillon, S.; Baranek, T.; Perreau, C.; Roughley, P. J.; Maquart, F. X.; Wegrowski, Y. Identification of Beta1 Integrin as Mediator of Melanoma Cell Adhesion to Lumican. Biochem. Biophys. Res. Commun. 2008, 365, 266−272. (392) Brezillon, S.; Radwanska, A.; Zeltz, C.; Malkowski, A.; Ploton, D.; Bobichon, H.; Perreau, C.; Malicka-Blaszkiewicz, M.; Maquart, F. X.; Wegrowski, Y. Lumican Core Protein Inhibits Melanoma Cell Migration Via Alterations of Focal Adhesion Complexes. Cancer Lett. 2009, 283, 92−100. (393) Radwanska, A.; Baczynska, D.; Nowak, D.; Brezillon, S.; Popow, A.; Maquart, F. X.; Wegrowski, Y.; Malicka-Blaszkiewicz, M. Lumican Affects Actin Cytoskeletal Organization in Human Melanoma A375 Cells. Life Sci. 2008, 83, 651−660. (394) Jeanne, A.; Untereiner, V.; Perreau, C.; Proult, I.; Gobinet, C.; Boulagnon-Rombi, C.; Terryn, C.; Martiny, L.; Brezillon, S.; Dedieu, S. Lumican Delays Melanoma Growth in Mice and Drives Tumor Molecular Assembly as Well as Response to Matrix-Targeted Tax2 Therapeutic Peptide. Sci. Rep. 2017, 7, 7700. (395) Zeltz, C.; Brezillon, S.; Kapyla, J.; Eble, J. A.; Bobichon, H.; Terryn, C.; Perreau, C.; Franz, C. M.; Heino, J.; Maquart, F. X.; et al. Lumican Inhibits Cell Migration through Alpha2beta1 Integrin. Exp. Cell Res. 2010, 316, 2922−2931. (396) Brezillon, S.; Venteo, L.; Ramont, L.; D’Onofrio, M. F.; Perreau, C.; Pluot, M.; Maquart, F. X.; Wegrowski, Y. Expression of Lumican, a Small Leucine-Rich Proteoglycan with Antitumour Activity, in Human Malignant Melanoma. Clin. Exp. Dermatol. 2007, 32, 405−416. (397) 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 Alpha2beta1 Receptor Activity and Downregulating Mmp-14 Expression. Thromb. Res. 2011, 128, 452−457. (398) 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. (399) Matsuda, Y.; Yamamoto, T.; Kudo, M.; Kawahara, K.; Kawamoto, M.; Nakajima, Y.; Koizumi, K.; Nakazawa, N.; Ishiwata, T.; Naito, Z. Expression and Roles of Lumican in Lung BL

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(416) Gesteira, T. F.; Coulson-Thomas, V. J.; Yuan, Y.; Zhang, J.; Nader, H. B.; Kao, W. W. Lumican Peptides: Rational Design Targeting Alk5/Tgfbri. Sci. Rep. 2017, 7, 42057. (417) Zeltz, C.; Brezillon, S.; Perreau, C.; Ramont, L.; Maquart, F. X.; Wegrowski, Y. Lumcorin: A Leucine-Rich Repeat 9-Derived Peptide from Human Lumican Inhibiting Melanoma Cell Migration. FEBS Lett. 2009, 583, 3027−3032. (418) Sulochana, K. N.; Fan, H.; Jois, S.; Subramanian, V.; Sun, F.; Kini, R. M.; Ge, R. Peptides Derived from Human Decorin LeucineRich Repeat 5 Inhibit Angiogenesis. J. Biol. Chem. 2005, 280, 27935− 27948. (419) Brassart-Pasco, S.; Senechal, K.; Thevenard, J.; Ramont, L.; Devy, J.; Di Stefano, L.; Dupont-Deshorgue, A.; Brezillon, S.; Feru, J.; Jazeron, J. F.; et al. Tetrastatin, the Nc1 Domain of the Alpha4(Iv) Collagen Chain: A Novel Potent Anti-Tumor Matrikine. PLoS One 2012, 7, e29587. (420) Thevenard, J.; Floquet, N.; Ramont, L.; Prost, E.; Nuzillard, J. M.; Dauchez, M.; Yezid, H.; Alix, A. J.; Maquart, F. X.; Monboisse, J. C.; et al. Structural and Antitumor Properties of the Ysnsg Cyclopeptide Derived from Tumstatin. Chem. Biol. 2006, 13, 1307− 1315. (421) Yi, M.; Ruoslahti, E. A Fibronectin Fragment Inhibits Tumor Growth, Angiogenesis, and Metastasis. Proc. Natl. Acad. Sci. U. S. A. 2001, 98, 620−624. (422) Guex, N.; Peitsch, M. C. Swiss-Model and the SwissPdbviewer: An Environment for Comparative Protein Modeling. Electrophoresis 1997, 18, 2714−2723. (423) Kajava, A. V. Structural Diversity of Leucine-Rich Repeat Proteins. J. Mol. Biol. 1998, 277, 519−527. (424) Dauvé, J. B. N.; Dauchez, M.; Brézillon, S.; Baud, S. Interactions between Lumican Derived Peptides and Mmp-14 Catalytic Domain Unraveled through a Combined Modeling and Statistical Approach. Submitted to BMC Bioinformatics, 2018. (425) Randles, M. J.; Humphries, M. J.; Lennon, R. Proteomic Definitions of Basement Membrane Composition in Health and Disease. Matrix Biol. 2017, 57−58, 12−28. (426) Cohen, I. R.; Grassel, S.; Murdoch, A. D.; Iozzo, R. V. Structural Characterization of the Complete Human Perlecan Gene and Its Promoter. Proc. Natl. Acad. Sci. U. S. A. 1993, 90, 10404− 10408. (427) Iozzo, R. V. Matrix Proteoglycans: From Molecular Design to Cellular Function. Annu. Rev. Biochem. 1998, 67, 609−652. (428) Sharma, B.; Iozzo, R. V. Transcriptional Silencing of Perlecan Gene Expression by Interferon-G. J. Biol. Chem. 1998, 273, 4642− 4646. (429) Iozzo, R. V.; Pillarisetti, J.; Sharma, B.; Murdoch, A. D.; Danielson, K. G.; Uitto, J.; Mauviel, A. Structural and Functional Characterization of the Human Perlecan Gene Promoter. Transcriptional Activation by Transforming Factor-B Via a Nuclear Factor 1Binding Element. J. Biol. Chem. 1997, 272, 5219−5228. (430) McCarthy, K. J. The Basement Membrane Proteoglycans Perlecan and Agrin: Something Old, Something New. Curr. Top. Membr. 2015, 76, 255−303. (431) Iozzo, R. V. Basement Membrane Proteoglycans: From Cellar to Ceiling. Nat. Rev. Mol. Cell Biol. 2005, 6, 646−656. (432) 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. (433) Wilusz, R. E.; Sanchez-Adams, J.; Guilak, F. The Structure and Function of the Pericellular Matrix of Articular Cartilage. Matrix Biol. 2014, 39, 25−32. (434) Hassell, J. R.; Yamada, Y.; Arikawa-Hirasawa, E. Role of Perlecan in Skeletal Development and Diseases. Glycoconjugate J. 2002, 19, 263−267. (435) Jochmann, K.; Bachvarova, V.; Vortkamp, A. Heparan Sulfate as a Regulator of Endochondral Ossification and Osteochondroma Development. Matrix Biol. 2014, 34, 55−63.

(436) Jung, M.; Lord, M. S.; Cheng, B.; Lyons, J. G.; Alkhouri, H.; Hughes, J. M.; McCarthy, S. J.; Iozzo, R. V.; Whitelock, J. M. Mast Cells Produce Novel Shorter Forms of Perlecan That Contain Functional Endorepellin: A Role in Angiogenesis and Wound Healing. J. Biol. Chem. 2013, 288, 3289−3304. (437) Hsueh, M. F.; Onnerfjord, P.; Kraus, V. B. Biomarkers and Proteomic Analysis of Osteoarthritis. Matrix Biol. 2014, 39, 56−66. (438) Christianson, H. C.; Belting, M. Heparan Sulfate Proteoglycan as a Cell-Surface Endocytosis Receptor. Matrix Biol. 2014, 35, 51−55. (439) Colombelli, C.; Palmisano, M.; Eshed-Eisenbach, Y.; Zambroni, D.; Pavoni, E.; Ferri, C.; Saccucci, S.; Nicole, S.; Soininen, R.; McKee, K. K.; et al. Perlecan Is Recruited by Dystroglycan to Nodes of Ranvier and Binds the Clustering Molecule Gliomedin. J. Cell Biol. 2015, 208, 313−329. (440) Fuki, I.; Iozzo, R. V.; Williams, K. J. Perlecan Heparan Sulfate Proteoglycan. A Novel Receptor That Mediates a Distinct Pathway for Ligand Catabolism. J. Biol. Chem. 2000, 275, 25742−25750. (441) Nugent, M. A.; Nugent, H. M.; Iozzo, R. V.; Sanchack, K.; Edelman, E. R. Perlecan Is Required to Inhibit Thrombosis after Deep Vascular Injury and Contributes to Endothelial Cell-Mediated Inhibition of Intimal Hyperplasia. Proc. Natl. Acad. Sci. U. S. A. 2000, 97, 6722−6727. (442) Zoeller, J. J.; McQuillan, A.; Whitelock, J.; Ho, S.-Y.; Iozzo, R. V. A Central Function for Perlecan in Skeletal Muscle and Cardiovascular Development. J. Cell Biol. 2008, 181, 381−394. (443) Zoeller, J. J.; Whitelock, J.; Iozzo, R. V. Perlecan Regulates Developmental Angiogenesis by Modulating the Vegf-Vegfr2 Axis. Matrix Biol. 2009, 28, 284−291. (444) Mathiak, M.; Yenisey, C.; Grant, D. S.; Sharma, B.; Iozzo, R. V. A Role for Perlecan in the Suppression of Growth and Invasion in Fibrosarcoma Cells. Cancer Res. 1997, 57, 2130−2136. (445) Sharma, B.; Handler, M.; Eichstetter, I.; Whitelock, J.; Nugent, M. A.; Iozzo, R. V. Antisense Targeting of Perlecan Blocks Tumor Growth and Angiogenesis in Vivo. J. Clin. Invest. 1998, 102, 1599− 1608. (446) Iozzo, R. V.; San Antonio, J. D. Heparan Sulfate Proteoglycans: Heavy Hitters in the Angiogenesis Arena. J. Clin. Invest. 2001, 108, 349−355. (447) Mongiat, M.; Sweeney, S.; 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. (448) 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. (449) Iozzo, R. V.; Zoeller, J. J.; Nystrom, A. Basement Membrane Proteoglycans: Modulators Par Excellence of Cancer Growth and Angiogenesis. Mol. Cells 2009, 27, 503−513. (450) Iozzo, R. V.; Sanderson, R. D. Proteoglycans in Cancer Biology, Tumour Microenvironment and Angiogenesis. J. Cell. Mol. Med. 2011, 15, 1013−1031. (451) Qiang, B.; Lim, S. Y.; Lekas, M.; Kuliszewski, M. A.; Wolff, R.; Osherov, A. B.; Rudenko, D.; Leong-Poi, H.; Noyan, H.; Husain, M.; et al. Perlecan Heparan Sulfate Proteoglycan Is a Critical Determinant of Angiogenesis in Response to Mouse Hind-Limb Ischemia. Can. J. Cardiol. 2014, 30, 1444−1451. (452) Poluzzi, C.; Iozzo, R. V.; Schaefer, L. Endostatin and Endorepellin: A Common Route of Action for Similar Angiostatic Cancer Avengers. Adv. Drug Delivery Rev. 2016, 97, 156−173. (453) 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. (454) Ning, L.; Xu, Z.; Furuya, N.; Nonaka, R.; Yamada, Y.; Arikawa-Hirasawa, E. Perlecan Inhibits Autophagy to Maintain Muscle Homeostasis in Mouse Soleus Muscle. Matrix Biol. 2015, 48, 26−35. (455) Willis, C. D.; Schaefer, L.; Iozzo, R. V. In Extracellular Matrix: Pathobiology and Signaling; Karamanos, N. K., Ed.; Walter de Gruyter GmbH & Co. KG: Berlin, 2012. BM

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(456) Peysselon, F.; Ricard-Blum, S. Heparin-Protein Interactions: From Affinity and Kinetics to Biological Roles. Application to an Interaction Network Regulating Angiogenesis. Matrix Biol. 2014, 35, 73−81. (457) Tanimoto, R.; Palladino, C.; Xu, S. Q.; Buraschi, S.; Neill, T.; Gomella, L. G.; Peiper, S. C.; Belfiore, A.; Iozzo, R. V.; Morrione, A. The Perlecan-Interacting Growth Factor Progranulin Regulates Ubiquitination, Sorting, and Lysosomal Degradation of Sortilin. Matrix Biol. 2017, 64, 27−39. (458) Neill, T.; Buraschi, S.; Goyal, A.; Sharpe, C.; Natkanski, E.; Schaefer, L.; Morrione, A.; Iozzo, R. V. Epha2 Is a Functional Receptor for the Growth Factor Progranulin. J. Cell Biol. 2016, 215, 687−703. (459) Ghadiali, R. S.; Guimond, S. E.; Turnbull, J. E.; Pisconti, A. Dynamic Changes in Heparan Sulfate During Muscle Differentiation and Ageing Regulate Myoblast Cell Fate and Fgf2 Signalling. Matrix Biol. 2017, 59, 54−68. (460) Patel, V. N.; Pineda, D. L.; Hoffman, M. P. The Function of Heparan Sulfate During Branching Morphogenesis. Matrix Biol. 2017, 57−58, 311−323. (461) Gonzalez, E. M.; Reed, C. C.; Bix, G.; Fu, J.; Zhang, Y.; Gopalakrishnan, B.; Greenspan, D. S.; Iozzo, R. V. Bmp-1/TolloidLike Metalloproteases Process Endorepellin, the Angiostatic CTerminal Fragment of Perlecan. J. Biol. Chem. 2005, 280, 7080−7087. (462) Cailhier, J.-F.; Sirois, I.; Raymond, M.-A.; Lepage, S.; Laplante, P.; Brassard, N.; Prat, A.; Iozzo, R. V.; Pshezhetsky, A. V.; Hebert, M.J. Caspase-3 Activation Triggers Extracellular Release of Cathepsin L and Endorepellin Proteolysis. J. Biol. Chem. 2008, 283, 27220−27229. (463) Le, B. V.; Kim, H.; Choi, J.; Kim, J.-H.; Hahn, M.-J.; Lee, C.; Kim, K. K.; Hwang, H.-Y. Crystal Structure of the Lg3 Domain of Endorepellin, an Angiogenesis Inhibitor. J. Mol. Biol. 2011, 414, 231− 242. (464) Parker, T. J.; Sampson, D. L.; Broszczak, D.; Chng, Y. L.; Carter, S. L.; Leavesley, D. I.; Parker, A. W.; Upton, Z. A Fragment of the Lg3 Peptide of Endorepellin Is Present in the Urine of Physically Active Mining Workers: A Potential Marker of Physical Activity. PLoS One 2012, 7, e33714. (465) Oda, O.; Shinzato, T.; Ohbayashi, K.; Takai, I.; Kunimatsu, M.; Maeda, K.; Yamanaka, N. Purification and Characterization of Perlecan Fragment in Urine of End-Stage Renal Failure Patients. Clin. Chim. Acta 1996, 255, 119−132. (466) Vuadens, F.; Benay, C.; Crettaz, D.; Gallot, D.; Sapin, V.; Schneider, P.; Bienvenut, W.-V.; Lemery, D.; Quadroni, M.; Dastugue, B.; et al. Identification of Biologic Markers of the Premature Rupture of Fetal Membranes: Proteomic Approach. Proteomics 2003, 3, 1521−1525. (467) O’Riordan, E.; Orlova, T. N.; Mendelev, N.; Patschan, D.; Kemp, R.; Chander, P. N.; Hu, R.; Hao, G.; Gross, S. S.; Iozzo, R. V.; et al. Urinary Proteomic Analysis of Chronic Renal Allograft Nephropathy. Proteomics: Clin. Appl. 2008, 2, 1025−1035. (468) Thadikkaran, L.; Crettaz, D.; Siegenthaler, M. A.; Gallot, D.; Sapin, V.; Iozzo, R. V.; Queloz, P. A.; Schneider, P.; Tissot, J. D. The Role of Proteomics in the Assessment of Premature Rupture of Fetal Membranes. Clin. Chim. Acta 2005, 360, 27−36. (469) Mauri, P.; Scarpa, A.; Nascimbeni, A. C.; Benazzi, L.; Parmagnani, E.; Mafficini, A.; Della Peruta, M.; Bassi, C.; Miyazaki, K.; Sorio, C. Identification of Proteins Released by Pancreatic Cancer Cells by Multidimensional Protein Identification Technology: A Strategy for Identification of Novel Cancer Markers. FASEB J. 2005, 19, 1125−1127. (470) Gronborg, M.; Kristiansen, T. Z.; Iwahori, A.; Chang, R.; Reddy, R.; Sato, N.; Molina, H.; Jensen, O. N.; Hruban, R. H.; Goggins, M. G.; et al. Biomarker Discovery from Pancreatic Cancer Secretome Using a Differential Proteomic Approach. Mol. Cell. Proteomics 2006, 5, 157−171. (471) Tsangaris, G. T.; Karamessinis, P.; Kolialexi, A.; Garbis, S. D.; Antsaklis, A.; Mavrou, A.; Fountoulakis, M. Proteomic Analysis of Amniotic Fluid in Pregnancies with Down Syndrome. Proteomics 2006, 6, 4410−4419.

(472) Aspinall-O’Dea, M.; Costello, E. The Pancreatic Cancer Proteome-Recent Advances and Future Promise. Proteomics: Clin. Appl. 2007, 1, 1066−1079. (473) Majek, P.; Reicheltova, Z.; Suttnar, J.; Cermak, J.; Dyr, J. E. Plasma Proteome Changes Associated with Refractory Cytopenia with Multilineage Dysplasia. Proteome Sci. 2011, 9, 64. (474) Surin, B.; Sachon, E.; Rougier, J.-P.; Steverlynck, C.; Garreau, C.; Lelongt, B.; Ronco, P.; Piedagnel, R. Lg3 Fragment of Endorepellin Is a Possible Biomarker of Severity in Lga Nephropathy. Proteomics 2013, 13, 142−152. (475) 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. Proteomics: Clin. Appl. 2008, 2, 23−32. (476) Bix, G.; Iozzo, R. V. Matrix Revolutions: ″Tails″ of BasementMembrane Components with Angiostatic Functions. Trends Cell Biol. 2005, 15, 52−60. (477) Bix, G.; Iozzo, R. V. Novel Interactions of Perlecan: Unraveling Perlecan’s Role in Angiogenesis. Microsc. Res. Tech. 2008, 71, 339−348. (478) 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 A2b1 Integrin and Vascular Endothelial Growth Factor Receptor 2 (Vegfr2). J. Biol. Chem. 2011, 286, 25947−25962. (479) Bix, G.; Fu, J.; Gonzalez, E.; 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 the A2b1 Integrin. J. Cell Biol. 2004, 166, 97−109. (480) Willis, C. D.; Poluzzi, C.; Mongiat, M.; Iozzo, R. V. Endorepellin Laminin-Like Globular Repeat 1/2 Domains Bind Ig3−5 of Vascular Endothelial Growth Factor(Vegf) Receptor 2 and Block Pro-Angiogenic Signaling by Vegfa in Endothelial Cells. FEBS J. 2013, 280, 2271−2294. (481) Bix, G.; Iozzo, R. A.; Woodall, B.; Burrows, M.; McQuillan, A.; Campbell, S.; Fields, G. B.; Iozzo, R. V. Endorepellin, the C-Terminal Angiostatic Module of Perlecan, Enhances Collagen-Platelet Responses Via the A2b1 Integrin Receptor. Blood 2007, 109, 3745− 3748. (482) Woodall, B. P.; Nystrom, A.; Iozzo, R. A.; Eble, J. A.; Niland, S.; Krieg, T.; Eckes, B.; Pozzi, A.; Iozzo, R. V. Integrin A2b1 Is the Required Receptor for Endorepellin Angiostatic Activity. J. Biol. Chem. 2008, 283, 2335−2343. (483) 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. (484) Goyal, A.; Poluzzi, C.; Willis, A. C.; Smythies, J.; Shellard, A.; Neill, T.; Iozzo, R. V. Endorepellin Affects Angiogenesis by Antagonizing Diverse Vegfr2- Evoked Signaling Pathways: Transcriptional Repression of Hif-1a and Vegfa and Concurrent Inhibition of Nfat1 Activation. J. Biol. Chem. 2012, 287, 43543−43556. (485) Bhattacharya, R.; Kwon, J.; Wang, E.; Mukherjee, P.; Mukhopadhyay, D. Src Homology 2 (Sh2) Domain Containing Protein Tyrosine Phosphatase-1 (Shp-1) Dephosphorylates Vegf Receptor-2 and Attenuates Endothelial DNA Synthesis, but Not Migration. J. Mol. Signaling 2014, 3, 8. (486) Choi, A. M. K.; Ryter, S. W.; Levine, B. Autophagy in Human Health and Disease. N. Engl. J. Med. 2013, 368, 651−662. (487) Goyal, A.; Gubbiotti, M. A.; Chery, D. R.; Han, L.; Iozzo, R. V. Endorepellin-Evoked Autophagy Contributes to Angiostasis. J. Biol. Chem. 2016, 291, 19245−19256. (488) Thoreen, C. C.; Kang, S. A.; Chang, J. W.; Liu, Q.; Zhang, J.; Gao, Y.; Reichling, L. J.; Sim, T.; Sabatini, D. M.; Gray, N. S. An AtpCompetitive Mammalian Target of Rapamycin Inhibitor Reveals Rapamycin-Resistant Functions of Mtorc1. J. Biol. Chem. 2009, 284, 8023−8032. BN

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(489) Heljasvaara, R.; Aikio, M.; Ruotsalainen, H.; Pihlajaniemi, T. Collagen Xviii in Tissue Homeostasis and Dysregulation - Lessons Learned from Model Organisms and Human Patients. Matrix Biol. 2017, 57−58, 55−75. (490) Nguyen, T. M. B.; Subramanian, I. V.; Xiao, X.; Ghosh, G.; Nguyen, P.; Kelekar, A.; Ramakrishnan, S. Endostatin Induces Autophagy in Endothelial Cells by Modulating Beclin 1 and BCatenin Levels. J. Cell. Mol. Med. 2009, 13, 3687−3698. (491) Fullgrabe, J.; Lynch-Day, M. A.; Heldring, N.; Li, W.; Struijik, R. B.; Ma, Q.; Hermanson, O.; Rosenfeld, M. G.; Klionsky, D. J.; Joseph, B. The Histone H4 Lysine Acetyltransferase Hmof Regulates the Outcome of Autophagy. Nature 2013, 500, 468−471. (492) Seok, S.; Fu, T.; Choi, S. E.; Li, Y.; Zhu, R.; Kumar, S.; Sun, X.; Yoon, G.; Kang, Y.; Zhong, W.; et al. Transcriptional Regulation of Autophagy by an Fxr-Creb Axis. Nature 2014, 516, 108−111. (493) Neill, T.; Andreuzzi, E.; Wang, Z.-X.; Peiper, S. C.; Mongiat, M.; Iozzo, R. B. Endorepellin Remodels the Endothelial Transcriptome toward a Pro-Autophagic and Pro-Mitophagic Gene Signature. J. Biol. Chem. 2018, 293, 12137. (494) Narendra, D. P.; Jin, S. M.; Tanaka, A.; Suen, D. F.; Gautier, C. A.; Shen, J.; Cookson, M. R.; Youle, R. J. Pink1 Is Selectively Stabilized on Impaired Mitochondria to Activate Parkin. PLoS Biol. 2010, 8, e1000298. (495) Narendra, D.; Tanaka, A.; Suen, D. F.; Youle, R. J. Parkin Is Recruited Selectively to Impaired Mitochondria and Promotes Their Autophagy. J. Cell Biol. 2008, 183, 795−803. (496) Cossarizza, A.; Baccarani-Contri, M.; Kalashnikova, G.; Franceschi, C. A New Method for the Cytofluorimetric Analysis of Mitochondrial Membrane Potential Using the J-Aggregate Forming Lipophilic Cation 5,5′,6,6′-Tetrachloro-1,1′,3,3′-Tetraethylbenzimidazolcarbocyanine Iodide (Jc-1). Biochem. Biophys. Res. Commun. 1993, 197, 40−45. (497) Bernfield, M.; Gotte, M.; Park, P. W.; Reizes, O.; Fitzgerald, M. L.; Lincecum, J.; Zako, M. Functions of Cell Surface Heparan Sulfate Proteoglycans. Annu. Rev. Biochem. 1999, 68, 729−777. (498) Afratis, N. A.; Nikitovic, D.; Multhaupt, H. A.; Theocharis, A. D.; Couchman, J. R.; Karamanos, N. K. Syndecans: Key Regulators of Cell Signaling and Biological Functions. FEBS J. 2017, 284, 27−41. (499) Piperigkou, Z.; Mohr, B.; Karamanos, N.; Gotte, M. Shed Proteoglycans in Tumor Stroma. Cell Tissue Res. 2016, 365, 643−655. (500) Couchman, J. R. Transmembrane Signaling Proteoglycans. Annu. Rev. Cell Dev. Biol. 2010, 26, 89−114. (501) Bourdon, M. A.; Krusius, T.; Campbell, S.; Schwartz, N. B.; Ruoslahti, E. Identification and Synthesis of a Recognition Signal for the Attachment of Glycosaminoglycans to Proteins. Proc. Natl. Acad. Sci. U. S. A. 1987, 84, 3194−3198. (502) Dews, I. C.; Mackenzie, K. R. Transmembrane Domains of the Syndecan Family of Growth Factor Coreceptors Display a Hierarchy of Homotypic and Heterotypic Interactions. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 20782−20787. (503) 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. (504) Lendorf, M. E.; Manon-Jensen, T.; Kronqvist, P.; Multhaupt, H. A.; Couchman, J. R. Syndecan-1 and Syndecan-4 Are Independent Indicators in Breast Carcinoma. J. Histochem. Cytochem. 2011, 59, 615−629. (505) Langford, J. K.; Stanley, M. J.; Cao, D.; Sanderson, R. D. Multiple Heparan Sulfate Chains Are Required for Optimal Syndecan1 Function. J. Biol. Chem. 1998, 273, 29965−29971. (506) Zhang, Y.; Wang, N.; Raab, R. W.; McKown, R. L.; Irwin, J. A.; Kwon, I.; van Kuppevelt, T. H.; Laurie, G. W. Targeting of Heparanase-Modified Syndecan-1 by Prosecretory Mitogen Lacritin Requires Conserved Core Gagal Plus Heparan and Chondroitin Sulfate as a Novel Hybrid Binding Site That Enhances Selectivity. J. Biol. Chem. 2013, 288, 12090−12101. (507) Deepa, S. S.; Yamada, S.; Zako, M.; Goldberger, O.; Sugahara, K. Chondroitin Sulfate Chains on Syndecan-1 and Syndecan-4 from

Normal Murine Mammary Gland Epithelial Cells Are Structurally and Functionally Distinct and Cooperate with Heparan Sulfate Chains to Bind Growth Factors. A Novel Function to Control Binding of Midkine, Pleiotrophin, and Basic Fibroblast Growth Factor. J. Biol. Chem. 2004, 279, 37368−37376. (508) Ramani, V. C.; Pruett, P. S.; Thompson, C. A.; DeLucas, L. D.; Sanderson, R. D. Heparan Sulfate Chains of Syndecan-1 Regulate Ectodomain Shedding. J. Biol. Chem. 2012, 287, 9952−9961. (509) Langford, J. K.; Yang, Y.; Kieber-Emmons, T.; Sanderson, R. D. Identification of an Invasion Regulatory Domain within the Core Protein of Syndecan-1. J. Biol. Chem. 2005, 280, 3467−3473. (510) Beauvais, D. M.; Jung, O.; Yang, Y.; Sanderson, R. D.; Rapraeger, A. C. Syndecan-1 (Cd138) Suppresses Apoptosis in Multiple Myeloma by Activating Igf1 Receptor: Prevention by Synstatinigf1r Inhibits Tumor Growth. Cancer Res. 2016, 76, 4981− 4993. (511) Whiteford, J. R.; Couchman, J. R. A Conserved Nxip Motif Is Required for Cell Adhesion Properties of the Syndecan-4 Ectodomain. J. Biol. Chem. 2006, 281, 32156−32163. (512) Kwon, M. J.; Park, J.; Jang, S.; Eom, C. Y.; Oh, E. S. The Conserved Phenylalanine in the Transmembrane Domain Enhances Heteromeric Interactions of Syndecans. J. Biol. Chem. 2016, 291, 872−881. (513) Altemeier, W. A.; Schlesinger, S. Y.; Buell, C. A.; Brauer, R.; Rapraeger, A. C.; Parks, W. C.; Chen, P. Transmembrane and Extracellular Domains of Syndecan-1 Have Distinct Functions in Regulating Lung Epithelial Migration and Adhesion. J. Biol. Chem. 2012, 287, 34927−34935. (514) Gopal, S.; Multhaupt, H. A. B.; Pocock, R.; Couchman, J. R. Cell-Extracellular Matrix and Cell-Cell Adhesion Are Linked by Syndecan-4. Matrix Biol. 2017, 60−61, 57−69. (515) Schulz, J. G.; Annaert, W.; Vandekerckhove, J.; Zimmermann, P.; De Strooper, B.; David, G. Syndecan 3 Intramembrane Proteolysis Is Presenilin/Gamma-Secretase-Dependent and Modulates Cytosolic Signaling. J. Biol. Chem. 2003, 278, 48651−48657. (516) Pasqualon, T.; Pruessmeyer, J.; Jankowski, V.; Babendreyer, A.; Groth, E.; Schumacher, J.; Koenen, A.; Weidenfeld, S.; Schwarz, N.; Denecke, B.; et al. A Cytoplasmic C-Terminal Fragment of Syndecan-1 Is Generated by Sequential Proteolysis and Antagonizes Syndecan-1 Dependent Lung Tumor Cell Migration. Oncotarget 2015, 6, 31295−31312. (517) Zong, F.; Fthenou, E.; Wolmer, N.; Hollosi, P.; Kovalszky, I.; Szilak, L.; Mogler, C.; Nilsonne, G.; Tzanakakis, G.; Dobra, K. Syndecan-1 and Fgf-2, but Not Fgf Receptor-1, Share a Common Transport Route and Co-Localize with Heparanase in the Nuclei of Mesenchymal Tumor Cells. PLoS One 2009, 4, e7346. (518) Gopal, S.; Sogaard, P.; Multhaupt, H. A.; Pataki, C.; Okina, E.; Xian, X.; Pedersen, M. E.; Stevens, T.; Griesbeck, O.; Park, P. W.; et al. Transmembrane Proteoglycans Control Stretch-Activated Channels to Set Cytosolic Calcium Levels. J. Cell Biol. 2015, 210, 1199−1211. (519) Jurjus, R. A.; Liu, Y.; Pal-Ghosh, S.; Tadvalkar, G.; Stepp, M. A. Primary Dermal Fibroblasts Derived from Sdc-1 Deficient Mice Migrate Faster and Have Altered Alphav Integrin Function. Wound Repair Regen. 2008, 16, 649−660. (520) Filmus, J.; Capurro, M. The Role of Glypicans in Hedgehog Signaling. Matrix Biol. 2014, 35, 248−252. (521) Capurro, M.; Martin, T.; Shi, W.; Filmus, J. Glypican-3 Binds to Frizzled and Plays a Direct Role in the Stimulation of Canonical Wnt Signaling. J. Cell Sci. 2014, 127, 1565−1575. (522) Nishiyama, A.; Dahlin, K. J.; Prince, J. T.; Johnstone, S. R.; Stallcup, W. B. The Primary Structure of Ng2, a Novel MembraneSpanning Proteoglycan. J. Cell Biol. 1991, 114, 359−371. (523) Lopez-Casillas, F.; Cheifetz, S.; Doody, J.; Andres, J. L.; Lane, W. S.; Massague, J. Structure and Expression of the Membrane Proteoglycan Betaglycan, a Component of the Tgf-Beta Receptor System. Cell 1991, 67, 785−795. (524) Elderbroom, J. L.; Huang, J. J.; Gatza, C. E.; Chen, J.; How, T.; Starr, M.; Nixon, A. B.; Blobe, G. C. Ectodomain Shedding of BO

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Tbetariii Is Required for Tbetariii-Mediated Suppression of Tgf-Beta Signaling and Breast Cancer Migration and Invasion. Mol. Biol. Cell 2014, 25, 2320−2332. (525) Reinhard, J.; Brosicke, N.; Theocharidis, U.; Faissner, A. The Extracellular Matrix Niche Microenvironment of Neural and Cancer Stem Cells in the Brain. Int. J. Biochem. Cell Biol. 2016, 81, 174−183. (526) Ibrahim, S. A.; Gadalla, R.; El-Ghonaimy, E. A.; Samir, O.; Mohamed, H. T.; Hassan, H.; Greve, B.; El-Shinawi, M.; Mohamed, M. M.; Gotte, M. Syndecan-1 Is a Novel Molecular Marker for Triple Negative Inflammatory Breast Cancer and Modulates the Cancer Stem Cell Phenotype Via the Il-6/Stat3, Notch and Egfr Signaling Pathways. Mol. Cancer 2017, 16, 57. (527) Bourguignon, L. Y.; Wong, G.; Earle, C.; Chen, L. Hyaluronan-Cd44v3 Interaction with Oct4-Sox2-Nanog Promotes Mir-302 Expression Leading to Self-Renewal, Clonal Formation, and Cisplatin Resistance in Cancer Stem Cells from Head and Neck Squamous Cell Carcinoma. J. Biol. Chem. 2012, 287, 32800−32824. (528) Rapraeger, A. C. Syndecan-Regulated Receptor Signaling. J. Cell Biol. 2000, 149, 995−998. (529) Rapraeger, A. C.; Krufka, A.; Olwin, B. B. Requirement of Heparan Sulfate for Bfgf-Mediated Fibroblast Growth and Myoblast Differentiation. Science 1991, 252, 1705−1708. (530) Gotte, M. Syndecans in Inflammation. FASEB J. 2003, 17, 575−591. (531) Kumar, A. V.; Katakam, S. K.; Urbanowitz, A. K.; Gotte, M. Heparan Sulphate as a Regulator of Leukocyte Recruitment in Inflammation. Curr. Protein Pept. Sci. 2015, 16, 77−86. (532) Monneau, Y. R.; Luo, L.; Sankaranarayanan, N. V.; Nagarajan, B.; Vives, R. R.; Baleux, F.; Desai, U. R.; Arenzana-Seidedos, F.; Lortat-Jacob, H. Solution Structure of Cxcl13 and Heparan Sulfate Binding Show That Gag Binding Site and Cellular Signalling Rely on Distinct Domains. Open Biol. 2017, 7, 170133. (533) Gallagher, J. Fell-Muir Lecture: Heparan Sulphate and the Art of Cell Regulation: A Polymer Chain Conducts the Protein Orchestra. Int. J. Exp. Pathol. 2015, 96, 203−231. (534) Connell, B. J.; Sadir, R.; Baleux, F.; Laguri, C.; Kleman, J. P.; Luo, L.; Arenzana-Seisdedos, F.; Lortat-Jacob, H. Heparan Sulfate Differentially Controls Cxcl12alpha- and Cxcl12gamma-Mediated Cell Migration through Differential Presentation to Their Receptor Cxcr4. Sci. Signaling 2016, 9, ra107. (535) Ramovs, V.; Te Molder, L.; Sonnenberg, A. The Opposing Roles of Laminin-Binding Integrins in Cancer. Matrix Biol. 2017, 57− 58, 213−243. (536) Hassan, H.; Greve, B.; Pavao, M. S.; Kiesel, L.; Ibrahim, S. A.; Gotte, M. Syndecan-1 Modulates Beta-Integrin-Dependent and Interleukin-6-Dependent Functions in Breast Cancer Cell Adhesion, Migration, and Resistance to Irradiation. FEBS J. 2013, 280, 2216− 2227. (537) Masouleh, B. K.; Ten Dam, G. B.; Wild, M. K.; Seelige, R.; van der Vlag, J.; Rops, A. L.; Echtermeyer, F. G.; Vestweber, D.; van Kuppevelt, T. H.; Kiesel, L.; et al. Role of the Heparan Sulfate Proteoglycan Syndecan-1 (Cd138) in Delayed-Type Hypersensitivity. J. Immunol. 2009, 182, 4985−4993. (538) Rapraeger, A. C. Synstatin: A Selective Inhibitor of the Syndecan-1-Coupled Igf1r-Alphavbeta3 Integrin Complex in Tumorigenesis and Angiogenesis. FEBS J. 2013, 280, 2207−2215. (539) Ibrahim, S. A.; Yip, G. W.; Stock, C.; Pan, J. W.; Neubauer, C.; Poeter, M.; Pupjalis, D.; Koo, C. Y.; Kelsch, R.; Schule, R.; et al. Targeting of Syndecan-1 by Microrna Mir-10b Promotes Breast Cancer Cell Motility and Invasiveness Via a Rho-Gtpase- and ECadherin-Dependent Mechanism. Int. J. Cancer 2012, 131, E884− 896. (540) Morgan, M. R.; Hamidi, H.; Bass, M. D.; Warwood, S.; Ballestrem, C.; Humphries, M. J. Syndecan-4 Phosphorylation Is a Control Point for Integrin Recycling. Dev. Cell 2013, 24, 472−485. (541) Whiteford, J. R.; Xian, X.; Chaussade, C.; Vanhaesebroeck, B.; Nourshargh, S.; Couchman, J. R. Syndecan-2 Is a Novel Ligand for the Protein Tyrosine Phosphatase Receptor Cd148. Mol. Biol. Cell 2011, 22, 3609−3624.

(542) Zen, K.; Liu, D. Q.; Li, L. M.; Chen, C. X.; Guo, Y. L.; Ha, B.; Chen, X.; Zhang, C. Y.; Liu, Y. The Heparan Sulfate Proteoglycan Form of Epithelial Cd44v3 Serves as a Cd11b/Cd18 CounterReceptor During Polymorphonuclear Leukocyte Transepithelial Migration. J. Biol. Chem. 2009, 284, 3768−3776. (543) Stallcup, W. B. Ng2 Proteoglycan Enhances Brain Tumor Progression by Promoting Beta-1 Integrin Activation in Both Cis and Trans Orientations. Cancers 2017, 9, 31. (544) Tabata, T.; Takei, Y. Morphogens, Their Identification and Regulation. Development 2004, 131, 703−712. (545) Alexander, C. M.; Reichsman, F.; Hinkes, M. T.; Lincecum, J.; Becker, K. A.; Cumberledge, S.; Bernfield, M. Syndecan-1 Is Required for Wnt-1-Induced Mammary Tumorigenesis in Mice. Nat. Genet. 2000, 25, 329−332. (546) Russell, J. O.; Monga, S. P. Wnt/Beta-Catenin Signaling in Liver Development, Homeostasis, and Pathobiology. Annu. Rev. Pathol.: Mech. Dis. 2018, 13, 351−378. (547) Ohkawara, B.; Glinka, A.; Niehrs, C. Rspo3 Binds Syndecan 4 and Induces Wnt/Pcp Signaling Via Clathrin-Mediated Endocytosis to Promote Morphogenesis. Dev. Cell 2011, 20, 303−314. (548) Astudillo, P.; Carrasco, H.; Larrain, J. Syndecan-4 Inhibits Wnt/Beta-Catenin Signaling through Regulation of Low-DensityLipoprotein Receptor-Related Protein (Lrp6) and R-Spondin 3. Int. J. Biochem. Cell Biol. 2014, 46, 103−112. (549) Zhang, Z.; Rankin, S. A.; Zorn, A. M. Syndecan4 Coordinates Wnt/Jnk and Bmp Signaling to Regulate Foregut Progenitor Development. Dev. Biol. 2016, 416, 187−199. (550) Xie, Z.; Khair, M.; Shaukat, I.; Netter, P.; Mainard, D.; Barre, L.; Ouzzine, M. Non-Canonical Wnt Induces Chondrocyte DeDifferentiation through Frizzled 6 and Dvl-2/B-Raf/Camkiialpha/ Syndecan 4 Axis. Cell Death Differ. 2018, DOI: 10.1038/s41418-0170050-y. (551) Alexander, C. M. The Wnt Signaling Landscape of Mammary Stem Cells and Breast Tumors. Prog. Mol. Biol. Transl. Sci. 2018, 153, 271−298. (552) Liu, B. Y.; McDermott, S. P.; Khwaja, S. S.; Alexander, C. M. The Transforming Activity of Wnt Effectors Correlates with Their Ability to Induce the Accumulation of Mammary Progenitor Cells. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 4158−4163. (553) Ibrahim, S. A.; Hassan, H.; Vilardo, L.; Kumar, S. K.; Kumar, A. V.; Kelsch, R.; Schneider, C.; Kiesel, L.; Eich, H. T.; Zucchi, I.; et al. Syndecan-1 (Cd138) Modulates Triple-Negative Breast Cancer Stem Cell Properties Via Regulation of Lrp-6 and Il-6-Mediated Stat3 Signaling. PLoS One 2013, 8, e85737. (554) Ren, Z.; van Andel, H.; de Lau, W.; Hartholt, R. B.; Maurice, M. M.; Clevers, H.; Kersten, M. J.; Spaargaren, M.; Pals, S. T. Syndecan-1 Promotes Wnt/Beta-Catenin Signaling in Multiple Myeloma by Presenting Wnts and R-Spondins. Blood 2018, 131, 982−994. (555) Manikowski, D.; Kastl, P.; Grobe, K. Taking the Occam’s Razor Approach to Hedgehog Lipidation and Its Role in Development. J. Dev. Biol. 2018, 6, 3. (556) Yin, Y.; Wang, A.; Feng, L.; Wang, Y.; Zhang, H.; Zhang, I.; Bany, B.; Ma, L. Heparan Sulfate Proteoglycan Sulfation Regulates Uterine Differentiation and Signaling During Embryo Implantation. Endocrinology 2018, 159, 2459. (557) Ortmann, C.; Pickhinke, U.; Exner, S.; Ohlig, S.; Lawrence, R.; Jboor, H.; Dreier, R.; Grobe, K. Sonic Hedgehog Processing and Release Are Regulated by Glypican Heparan Sulfate Proteoglycans. J. Cell Sci. 2015, 128, 2374−2385. (558) Capurro, M.; Shi, W.; Izumikawa, T.; Kitagawa, H.; Filmus, J. Processing by Convertases Is Required for Glypican-3-Induced Inhibition of Hedgehog Signaling. J. Biol. Chem. 2015, 290, 7576− 7585. (559) Capurro, M.; Izumikawa, T.; Suarez, P.; Shi, W.; Cydzik, M.; Kaneiwa, T.; Gariepy, J.; Bonafe, L.; Filmus, J. Glypican-6 Promotes the Growth of Developing Long Bones by Stimulating Hedgehog Signaling. J. Cell Biol. 2017, 216, 2911−2926. BP

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(560) Kamimura, K.; Rhodes, J. M.; Ueda, R.; McNeely, M.; Shukla, D.; Kimata, K.; Spear, P. G.; Shworak, N. W.; Nakato, H. Regulation of Notch Signaling by Drosophila Heparan Sulfate 3-O Sulfotransferase. J. Cell Biol. 2004, 166, 1069−1079. (561) Pisconti, A.; Cornelison, D. D.; Olguin, H. C.; Antwine, T. L.; Olwin, B. B. Syndecan-3 and Notch Cooperate in Regulating Adult Myogenesis. J. Cell Biol. 2010, 190, 427−441. (562) Aster, J. C.; Pear, W. S.; Blacklow, S. C. The Varied Roles of Notch in Cancer. Annu. Rev. Pathol.: Mech. Dis. 2017, 12, 245−275. (563) Zhao, N.; Liu, H.; Lilly, B. Reciprocal Regulation of Syndecan2 and Notch Signaling in Vascular Smooth Muscle Cells. J. Biol. Chem. 2012, 287, 16111−16120. (564) Xu, D.; Hu, J.; Xu, S.; De Bruyne, E.; Menu, E.; Van Camp, B.; Vanderkerken, K.; Van Valckenborgh, E. Dll1/Notch Activation Accelerates Multiple Myeloma Disease Development by Promoting Cd138+ Mm-Cell Proliferation. Leukemia 2012, 26, 1402−1405. (565) Nikolova, V.; Koo, C. Y.; Ibrahim, S. A.; Wang, Z.; Spillmann, D.; Dreier, R.; Kelsch, R.; Fischgrabe, J.; Smollich, M.; Rossi, L. H.; et al. Differential Roles for Membrane-Bound and Soluble Syndecan-1 (Cd138) in Breast Cancer Progression. Carcinogenesis 2009, 30, 397− 407. (566) Elenius, V.; Gotte, M.; Reizes, O.; Elenius, K.; Bernfield, M. Inhibition by the Soluble Syndecan-1 Ectodomains Delays Wound Repair in Mice Overexpressing Syndecan-1. J. Biol. Chem. 2004, 279, 41928−41935. (567) Nam, E. J.; Park, P. W. Shedding of Cell Membrane-Bound Proteoglycans. Methods Mol. Biol. 2012, 836, 291−305. (568) Traister, A.; Shi, W.; Filmus, J. Mammalian Notum Induces the Release of Glypicans and Other Gpi-Anchored Proteins from the Cell Surface. Biochem. J. 2008, 410, 503−511. (569) Ayers, K. L.; Mteirek, R.; Cervantes, A.; Lavenant-Staccini, L.; Therond, P. P.; Gallet, A. Dally and Notum Regulate the Switch between Low and High Level Hedgehog Pathway Signalling. Development 2012, 139, 3168−3179. (570) Giraldez, A. J.; Copley, R. R.; Cohen, S. M. Hspg Modification by the Secreted Enzyme Notum Shapes the Wingless Morphogen Gradient. Dev. Cell 2002, 2, 667−676. (571) Kovalszky, I.; Hjerpe, A.; Dobra, K. Nuclear Translocation of Heparan Sulfate Proteoglycans and Their Functional Significance. Biochim. Biophys. Acta, Gen. Subj. 2014, 1840, 2491−2497. (572) Purushothaman, A.; Hurst, D. R.; Pisano, C.; Mizumoto, S.; Sugahara, K.; Sanderson, R. D. Heparanase-Mediated Loss of Nuclear Syndecan-1 Enhances Histone Acetyltransferase (Hat) Activity to Promote Expression of Genes That Drive an Aggressive Tumor Phenotype. J. Biol. Chem. 2011, 286, 30377−30383. (573) Kong, X.; Qian, J.; Chen, L. S.; Wang, Y. C.; Wang, J. L.; Chen, H.; Weng, Y. R.; Zhao, S. L.; Hong, J.; Chen, Y. X.; et al. Synbindin in Extracellular Signal-Regulated Protein Kinase Spatial Regulation and Gastric Cancer Aggressiveness. J. Natl. Cancer Inst. 2013, 105, 1738−1749. (574) Bourguignon, L. Y.; Gilad, E.; Brightman, A.; Diedrich, F.; Singleton, P. Hyaluronan-Cd44 Interaction with Leukemia-Associated Rhogef and Epidermal Growth Factor Receptor Promotes Rho/Ras Co-Activation, Phospholipase C Epsilon-Ca2+ Signaling, and Cytoskeleton Modification in Head and Neck Squamous Cell Carcinoma Cells. J. Biol. Chem. 2006, 281, 14026−14040. (575) Chatterjee, N.; Stegmuller, J.; Schatzle, P.; Karram, K.; Koroll, M.; Werner, H. B.; Nave, K. A.; Trotter, J. Interaction of Syntenin-1 and the Ng2 Proteoglycan in Migratory Oligodendrocyte Precursor Cells. J. Biol. Chem. 2008, 283, 8310−8317. (576) Fares, J.; Kashyap, R.; Zimmermann, P. Syntenin: Key Player in Cancer Exosome Biogenesis and Uptake? Cell. Adh. Migr. 2017, 11, 124−126. (577) Zimmermann, P.; Zhang, Z.; Degeest, G.; Mortier, E.; Leenaerts, I.; Coomans, C.; Schulz, J.; N’Kuli, F.; Courtoy, P. J.; David, G. Syndecan Recycling [Corrected] Is Controlled by SynteninPip2 Interaction and Arf6. Dev. Cell 2005, 9, 377−388. (578) Koo, B. K.; Jung, Y. S.; Shin, J.; Han, I.; Mortier, E.; Zimmermann, P.; Whiteford, J. R.; Couchman, J. R.; Oh, E. S.; Lee,

W. Structural Basis of Syndecan-4 Phosphorylation as a Molecular Switch to Regulate Signaling. J. Mol. Biol. 2006, 355, 651−663. (579) Benassi, M. S.; Pazzaglia, L.; Chiechi, A.; Alberghini, M.; Conti, A.; Cattaruzza, S.; Wassermann, B.; Picci, P.; Perris, R. Ng2 Expression Predicts the Metastasis Formation in Soft-Tissue Sarcoma Patients. J. Orthop. Res. 2009, 27, 135−140. (580) Farnedi, A.; Rossi, S.; Bertani, N.; Gulli, M.; Silini, E. M.; Mucignat, M. T.; Poli, T.; Sesenna, E.; Lanfranco, D.; Montebugnoli, L.; et al. Proteoglycan-Based Diversification of Disease Outcome in Head and Neck Cancer Patients Identifies Ng2/Cspg4 and Syndecan2 as Unique Relapse and Overall Survival Predicting Factors. BMC Cancer 2015, 15, 352. (581) Hsu, N. C.; Nien, P. Y.; Yokoyama, K. K.; Chu, P. Y.; Hou, M. F. High Chondroitin Sulfate Proteoglycan 4 Expression Correlates with Poor Outcome in Patients with Breast Cancer. Biochem. Biophys. Res. Commun. 2013, 441, 514−518. (582) Jensen, K. B.; Jones, J.; Watt, F. M. A Stem Cell Gene Expression Profile of Human Squamous Cell Carcinomas. Cancer Lett. 2008, 272, 23−31. (583) Keleg, S.; Titov, A.; Heller, A.; Giese, T.; Tjaden, C.; Ahmad, S. S.; Gaida, M. M.; Bauer, A. S.; Werner, J.; Giese, N. A. Chondroitin Sulfate Proteoglycan Cspg4 as a Novel Hypoxia-Sensitive Marker in Pancreatic Tumors. PLoS One 2014, 9, e100178. (584) Nicolosi, P. A.; Dallatomasina, A.; Perris, R. Theranostic Impact of Ng2/Cspg4 Proteoglycan in Cancer. Theranostics 2015, 5, 530−544. (585) Rivera, Z.; Ferrone, S.; Wang, X.; Jube, S.; Yang, H.; Pass, H. I.; Kanodia, S.; Gaudino, G.; Carbone, M. Cspg4 as a Target of Antibody-Based Immunotherapy for Malignant Mesothelioma. Clin. Cancer Res. 2012, 18, 5352−5363. (586) Schoenfeld, A. J.; Wang, X.; Wang, Y.; Hornicek, F. J.; Nielsen, G. P.; Duan, Z.; Ferrone, S.; Schwab, J. H. Cspg4 as a Prognostic Biomarker in Chordoma. Spine J. 2016, 16, 722−727. (587) Svendsen, A.; Verhoeff, J. J.; Immervoll, H.; Brogger, J. C.; Kmiecik, J.; Poli, A.; Netland, I. A.; Prestegarden, L.; Planaguma, J.; Torsvik, A.; et al. Expression of the Progenitor Marker Ng2/Cspg4 Predicts Poor Survival and Resistance to Ionising Radiation in Glioblastoma. Acta Neuropathol. 2011, 122, 495−510. (588) Wang, X.; Osada, T.; Wang, Y.; Yu, L.; Sakakura, K.; Katayama, A.; McCarthy, J. B.; Brufsky, A.; Chivukula, M.; Khoury, T.; et al. Cspg4 Protein as a New Target for the Antibody-Based Immunotherapy of Triple-Negative Breast Cancer. J. Natl. Cancer Inst. 2010, 102, 1496−1512. (589) Westekemper, H.; Karimi, S.; Susskind, D.; Anastassiou, G.; Freistuhler, M.; Meller, D.; Zeschnigk, M.; Steuhl, K. P.; Bornfeld, N.; Schmid, K. W.; et al. Expression of Mcsp and Prame in Conjunctival Melanoma. Br. J. Ophthalmol. 2010, 94, 1322−1327. (590) Zangrando, A.; Intini, F.; te Kronnie, G.; Basso, G. Validation of Ng2 Antigen in Identifying Bp-All Patients with Mll Rearrangements Using Qualitative and Quantitative Flow Cytometry: A Prospective Study. Leukemia 2008, 22, 858−861. (591) Cheever, M. A.; Allison, J. P.; Ferris, A. S.; Finn, O. J.; Hastings, B. M.; Hecht, T. T.; Mellman, I.; Prindiville, S. A.; Viner, J. L.; Weiner, L. M.; et al. The Prioritization of Cancer Antigens: A National Cancer Institute Pilot Project for the Acceleration of Translational Research. Clin. Cancer Res. 2009, 15, 5323−5337. (592) Hafner, C.; Breiteneder, H.; Ferrone, S.; Thallinger, C.; Wagner, S.; Schmidt, W. M.; Jasinska, J.; Kundi, M.; Wolff, K.; Zielinski, C. C.; et al. Suppression of Human Melanoma Tumor Growth in Scid Mice by a Human High Molecular Weight-Melanoma Associated Antigen (Hmw-Maa) Specific Monoclonal Antibody. Int. J. Cancer 2005, 114, 426−432. (593) de Bruyn, M.; Rybczynska, A. A.; Wei, Y.; Schwenkert, M.; Fey, G. H.; Dierckx, R. A.; van Waarde, A.; Helfrich, W.; Bremer, E. Melanoma-Associated Chondroitin Sulfate Proteoglycan (Mcsp)Targeted Delivery of Soluble Trail Potently Inhibits Melanoma Outgrowth in Vitro and in Vivo. Mol. Cancer 2010, 9, 301. (594) Amoury, M.; Mladenov, R.; Nachreiner, T.; Pham, A. T.; Hristodorov, D.; Di Fiore, S.; Helfrich, W.; Pardo, A.; Fey, G.; BQ

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Schwenkert, M.; et al. A Novel Approach for Targeted Elimination of Cspg4-Positive Triple-Negative Breast Cancer Cells Using a Map Tau-Based Fusion Protein. Int. J. Cancer 2016, 139, 916−927. (595) Brehm, H.; Niesen, J.; Mladenov, R.; Stein, C.; Pardo, A.; Fey, G.; Helfrich, W.; Fischer, R.; Gattenlohner, S.; Barth, S. A Cspg4Specific Immunotoxin Kills Rhabdomyosarcoma Cells and Binds to Primary Tumor Tissues. Cancer Lett. 2014, 352, 228−235. (596) Eng, M. S.; Kaur, J.; Prasmickaite, L.; Engesaeter, B. O.; Weyergang, A.; Skarpen, E.; Berg, K.; Rosenblum, M. G.; Maelandsmo, G. M.; Hogset, A.et al. Enhanced Targeting of TripleNegative Breast Carcinoma and Malignant Melanoma by Photochemical Internalization of Cspg4-Targeting Immunotoxins. Photochem. Photobiol. Sci. 2018, 17, 539. (597) Tordsson, J. M.; Ohlsson, L. G.; Abrahmsen, L. B.; Karlstrom, P. J.; Lando, P. A.; Brodin, T. N. Phage-Selected Primate Antibodies Fused to Superantigens for Immunotherapy of Malignant Melanoma. Cancer Immunol. Immunother. 2000, 48, 691−702. (598) Geldres, C.; Savoldo, B.; Hoyos, V.; Caruana, I.; Zhang, M.; Yvon, E.; Del Vecchio, M.; Creighton, C. J.; Ittmann, M.; Ferrone, S.; et al. T Lymphocytes Redirected against the Chondroitin Sulfate Proteoglycan-4 Control the Growth of Multiple Solid Tumors Both in Vitro and in Vivo. Clin. Cancer Res. 2014, 20, 962−971. (599) Torisu-Itakura, H.; Schoellhammer, H. F.; Sim, M. S.; Irie, R. F.; Hausmann, S.; Raum, T.; Baeuerle, P. A.; Morton, D. L. Redirected Lysis of Human Melanoma Cells by a Mcsp/Cd3Bispecific Bite Antibody That Engages Patient-Derived T Cells. J. Immunother. 2011, 34, 597−605. (600) Li, F.; Zhang, T.; Cao, L.; Zhang, Y. Chimeric Antigen Receptor T Cell Based Immunotherapy for Cancer. Curr. Stem Cell Res. Ther. 2018, 13, 327−335. (601) 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. Invest. 2008, 118, 89−99. (602) Hara, H.; Takahashi, T.; Serada, S.; Fujimoto, M.; Ohkawara, T.; Nakatsuka, R.; Harada, E.; Nishigaki, T.; Takahashi, Y.; Nojima, S.; et al. Overexpression of Glypican-1 Implicates Poor Prognosis and Their Chemoresistance in Oesophageal Squamous Cell Carcinoma. Br. J. Cancer 2016, 115, 66−75. (603) Li, J.; Chen, Y.; Guo, X.; Zhou, L.; Jia, Z.; Peng, Z.; Tang, Y.; Liu, W.; Zhu, B.; Wang, L.; et al. Gpc1 Exosome and Its Regulatory Mirnas Are Specific Markers for the Detection and Target Therapy of Colorectal Cancer. J. Cell Mol. Med. 2017, 21, 838−847. (604) Li, J.; Li, B.; Ren, C.; Chen, Y.; Guo, X.; Zhou, L.; Peng, Z.; Tang, Y.; Chen, Y.; Liu, W.; et al. The Clinical Significance of Circulating Gpc1 Positive Exosomes and Its Regulative Mirnas in Colon Cancer Patients. Oncotarget 2017, 8, 101189−101202. (605) Lu, H.; Niu, F.; Liu, F.; Gao, J.; Sun, Y.; Zhao, X. Elevated Glypican-1 Expression Is Associated with an Unfavorable Prognosis in Pancreatic Ductal Adenocarcinoma. Cancer Med. 2017, 6, 1181− 1191. (606) Matsuzaki, S.; Serada, S.; Hiramatsu, K.; Nojima, S.; Matsuzaki, S.; Ueda, Y.; Ohkawara, T.; Mabuchi, S.; Fujimoto, M.; Morii, E.; et al. Anti-Glypican-1 Antibody-Drug Conjugate Exhibits Potent Preclinical Antitumor Activity against Glypican-1 Positive Uterine Cervical Cancer. Int. J. Cancer 2018, 142, 1056−1066. (607) Melo, S. A.; Luecke, L. B.; Kahlert, C.; Fernandez, A. F.; Gammon, S. T.; Kaye, J.; LeBleu, V. S.; Mittendorf, E. A.; Weitz, J.; Rahbari, N.; et al. Glypican-1 Identifies Cancer Exosomes and Detects Early Pancreatic Cancer. Nature 2015, 523, 177−182. (608) Su, G.; Meyer, K.; Nandini, C. D.; Qiao, D.; Salamat, S.; Friedl, A. Glypican-1 Is Frequently Overexpressed in Human Gliomas and Enhances Fgf-2 Signaling in Glioma Cells. Am. J. Pathol. 2006, 168, 2014−2026. (609) Saito, T.; Sugiyama, K.; Hama, S.; Yamasaki, F.; Takayasu, T.; Nosaka, R.; Onishi, S.; Muragaki, Y.; Kawamata, T.; Kurisu, K. High Expression of Glypican-1 Predicts Dissemination and Poor Prognosis in Glioblastomas. World Neurosurg. 2017, 105, 282−288.

(610) Frampton, A. E.; Prado, M. M.; Lopez-Jimenez, E.; FajardoPuerta, A. B.; Jawad, Z. A. R.; Lawton, P.; Giovannetti, E.; Habib, N. A.; Castellano, L.; Stebbing, J.; et al. Glypican-1 Is Enriched in Circulating-Exosomes in Pancreatic Cancer and Correlates with Tumor Burden. Oncotarget 2018, 9, 19006−19013. (611) Lewis, J. M.; Vyas, A. D.; Qiu, Y.; Messer, K. S.; White, R.; Heller, M. J. Integrated Analysis of Exosomal Protein Biomarkers on Alternating Current Electrokinetic Chips Enables Rapid Detection of Pancreatic Cancer in Patient Blood. ACS Nano 2018, 12, 3311−3320. (612) Hu, J.; Sheng, Y.; Kwak, K. J.; Shi, J.; Yu, B.; Lee, L. J. A Signal-Amplifiable Biochip Quantifies Extracellular Vesicle-Associated Rnas for Early Cancer Detection. Nat. Commun. 2017, 8, 1683. (613) Li, N.; Fu, H.; Hewitt, S. M.; Dimitrov, D. S.; Ho, M. Therapeutically Targeting Glypican-2 Via Single-Domain AntibodyBased Chimeric Antigen Receptors and Immunotoxins in Neuroblastoma. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, E6623−E6631. (614) Bosse, K. R.; Raman, P.; Zhu, Z.; Lane, M.; Martinez, D.; Heitzeneder, S.; Rathi, K. S.; Kendsersky, N. M.; Randall, M.; Donovan, L.; et al. Identification of Gpc2 as an Oncoprotein and Candidate Immunotherapeutic Target in High-Risk Neuroblastoma. Cancer Cell 2017, 32, 295−309 e212. (615) Xiang, Y. Y.; Ladeda, V.; Filmus, J. Glypican-3 Expression Is Silenced in Human Breast Cancer. Oncogene 2001, 20, 7408−7412. (616) Zhu, Z.; Friess, H.; Kleeff, J.; Wang, L.; Wirtz, M.; Zimmermann, A.; Korc, M.; Buchler, M. W. Glypican-3 Expression Is Markedly Decreased in Human Gastric Cancer but Not in Esophageal Cancer. Am. J. Surg. 2002, 184, 78−83. (617) Valsechi, M. C.; Oliveira, A. B.; Conceicao, A. L.; Stuqui, B.; Candido, N. M.; Provazzi, P. J.; de Araujo, L. F.; Silva, W. A., Jr.; de Freitas Calmon, M.; Rahal, P. Gpc3 Reduces Cell Proliferation in Renal Carcinoma Cell Lines. BMC Cancer 2014, 14, 631. (618) Murthy, S. S.; Shen, T.; De Rienzo, A.; Lee, W. C.; Ferriola, P. C.; Jhanwar, S. C.; Mossman, B. T.; Filmus, J.; Testa, J. R. Expression of Gpc3, an X-Linked Recessive Overgrowth Gene, Is Silenced in Malignant Mesothelioma. Oncogene 2000, 19, 410−416. (619) Tretiakova, M.; Zynger, D. L.; Luan, C.; Andeen, N. K.; Finn, L. S.; Kocherginsky, M.; Teh, B. T.; Yang, X. J. Glypican 3 Overexpression in Primary and Metastatic Wilms Tumors. Virchows Arch. 2015, 466, 67−76. (620) Nakatsura, T.; Kageshita, T.; Ito, S.; Wakamatsu, K.; Monji, M.; Ikuta, Y.; Senju, S.; Ono, T.; Nishimura, Y. Identification of Glypican-3 as a Novel Tumor Marker for Melanoma. Clin. Cancer Res. 2004, 10, 6612−6621. (621) Yamanaka, K.; Ito, Y.; Okuyama, N.; Noda, K.; Matsumoto, H.; Yoshida, H.; Miyauchi, A.; Capurro, M.; Filmus, J.; Miyoshi, E. Immunohistochemical Study of Glypican 3 in Thyroid Cancer. Oncology 2007, 73, 389−394. (622) Lin, Q.; Xiong, L. W.; Pan, X. F.; Gen, J. F.; Bao, G. L.; Sha, H. F.; Feng, J. X.; Ji, C. Y.; Chen, M. Expression of Gpc3 Protein and Its Significance in Lung Squamous Cell Carcinoma. Med. Oncol. 2012, 29, 663−669. (623) Yu, X.; Li, Y.; Chen, S. W.; Shi, Y.; Xu, F. Differential Expression of Glypican-3 (Gpc3) in Lung Squamous Cell Carcinoma and Lung Adenocarcinoma and Its Clinical Significance. GMR, Genet. Mol. Res. 2015, 14, 10185−10192. (624) Capurro, M.; Wanless, I. R.; Sherman, M.; Deboer, G.; Shi, W.; Miyoshi, E.; Filmus, J. Glypican-3: A Novel Serum and Histochemical Marker for Hepatocellular Carcinoma. Gastroenterology 2003, 125, 89−97. (625) Chen, I. P.; Ariizumi, S.; Nakano, M.; Yamamoto, M. Positive Glypican-3 Expression in Early Hepatocellular Carcinoma Predicts Recurrence after Hepatectomy. J. Gastroenterol. 2014, 49, 117−125. (626) Wang, L.; Pan, L.; Yao, M.; Cai, Y.; Dong, Z.; Yao, D. Expression of Oncofetal Antigen Glypican-3 Associates Significantly with Poor Prognosis in Hbv-Related Hepatocellular Carcinoma. Oncotarget 2016, 7, 42150−42158. (627) Ofuji, K.; Saito, K.; Suzuki, S.; Shimomura, M.; Shirakawa, H.; Nobuoka, D.; Sawada, Y.; Yoshimura, M.; Tsuchiya, N.; Takahashi, M.; et al. Perioperative Plasma Glypican-3 Level May Enable BR

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Prediction of the Risk of Recurrence after Surgery in Patients with Stage I Hepatocellular Carcinoma. Oncotarget 2017, 8, 37835−37844. (628) Suzuki, S.; Sakata, J.; Utsumi, F.; Sekiya, R.; Kajiyama, H.; Shibata, K.; Kikkawa, F.; Nakatsura, T. Efficacy of Glypican-3-Derived Peptide Vaccine Therapy on the Survival of Patients with Refractory Ovarian Clear Cell Carcinoma. Oncoimmunology 2016, 5, e1238542. (629) Liu, H.; Yang, C.; Lu, W.; Zeng, Y. Prognostic Significance of Glypican-3 Expression in Hepatocellular Carcinoma: A Meta-Analysis. Medicine (Philadelphia, PA, U. S.) 2018, 97, e9702. (630) Zhang, Y.; Wang, J.; Dong, F.; Li, H.; Hou, Y. The Role of Gpc5 in Lung Metastasis of Salivary Adenoid Cystic Carcinoma. Arch. Oral Biol. 2014, 59, 1172−1182. (631) Williamson, D.; Selfe, J.; Gordon, T.; Lu, Y. J.; PritchardJones, K.; Murai, K.; Jones, P.; Workman, P.; Shipley, J. Role for Amplification and Expression of Glypican-5 in Rhabdomyosarcoma. Cancer Res. 2007, 67, 57−65. (632) Li, F.; Shi, W.; Capurro, M.; Filmus, J. Glypican-5 Stimulates Rhabdomyosarcoma Cell Proliferation by Activating Hedgehog Signaling. J. Cell Biol. 2011, 192, 691−704. (633) Liu, T.; Zhang, X.; Sha, K.; Liu, X.; Zhang, L.; Wang, B. Mir709 up-Regulated in Hepatocellular Carcinoma, Promotes Proliferation and Invasion by Targeting Gpc5. Cell Proliferation 2015, 48, 330−337. (634) Zhang, C.; Liu, Z.; Wang, L.; Qiao, B.; Du, E.; Li, L.; Xu, Y.; Zhang, Z. Prognostic Significance of Gpc5 Expression in Patients with Prostate Cancer. Tumor Biol. 2016, 37, 6413−6418. (635) Yang, X.; Zhang, Z.; Qiu, M.; Hu, J.; Fan, X.; Wang, J.; Xu, L.; Yin, R. Glypican-5 Is a Novel Metastasis Suppressor Gene in NonSmall Cell Lung Cancer. Cancer Lett. 2013, 341, 265−273. (636) Yuan, S.; Yu, Z.; Liu, Q.; Zhang, M.; Xiang, Y.; Wu, N.; Wu, L.; Hu, Z.; Xu, B.; Cai, T.; et al. Gpc5, a Novel Epigenetically Silenced Tumor Suppressor, Inhibits Tumor Growth by Suppressing Wnt/ Beta-Catenin Signaling in Lung Adenocarcinoma. Oncogene 2016, 35, 6120−6131. (637) Li, Y.; Sheu, C. C.; Ye, Y.; de Andrade, M.; Wang, L.; Chang, S. C.; Aubry, M. C.; Aakre, J. A.; Allen, M. S.; Chen, F.; et al. Genetic Variants and Risk of Lung Cancer in Never Smokers: A GenomeWide Association Study. Lancet Oncol. 2010, 11, 321−330. (638) Harada, E.; Serada, S.; Fujimoto, M.; Takahashi, Y.; Takahashi, T.; Hara, H.; Nakatsuka, R.; Sugase, T.; Nishigaki, T.; Saito, Y.; et al. Glypican-1 Targeted Antibody-Based Therapy Induces Preclinical Antitumor Activity against Esophageal Squamous Cell Carcinoma. Oncotarget 2017, 8, 24741−24752. (639) Zhang, Y. F.; Ho, M. Humanization of High-Affinity Antibodies Targeting Glypican-3 in Hepatocellular Carcinoma. Sci. Rep. 2016, 6, 33878. (640) Gao, W.; Tang, Z.; Zhang, Y. F.; Feng, M.; Qian, M.; Dimitrov, D. S.; Ho, M. Immunotoxin Targeting Glypican-3 Regresses Liver Cancer Via Dual Inhibition of Wnt Signalling and Protein Synthesis. Nat. Commun. 2015, 6, 6536. (641) Fleming, B. D.; Ho, M. Glypican-3 Targeting Immunotoxins for the Treatment of Liver Cancer. Toxins 2016, 8, 274. (642) Bi, Y.; Jiang, H.; Wang, P.; Song, B.; Wang, H.; Kong, X.; Li, Z. Treatment of Hepatocellular Carcinoma with a Gpc3-Targeted Bispecific T Cell Engager. Oncotarget 2017, 8, 52866−52876. (643) Ishiguro, T.; Sano, Y.; Komatsu, S. I.; Kamata-Sakurai, M.; Kaneko, A.; Kinoshita, Y.; Shiraiwa, H.; Azuma, Y.; Tsunenari, T.; Kayukawa, Y.et al. An Anti-Glypican 3/Cd3 Bispecific T CellRedirecting Antibody for Treatment of Solid Tumors. Sci. Transl. Med. 2017, 9, eaal4291. (644) Jiang, Z.; Jiang, X.; Chen, S.; Lai, Y.; Wei, X.; Li, B.; Lin, S.; Wang, S.; Wu, Q.; Liang, Q.; et al. Anti-Gpc3-Car T Cells Suppress the Growth of Tumor Cells in Patient-Derived Xenografts of Hepatocellular Carcinoma. Front. Immunol. 2017, 7, 690. (645) Yu, M.; Luo, H.; Fan, M.; Wu, X.; Shi, B.; Di, S.; Liu, Y.; Pan, Z.; Jiang, H.; Li, Z. Development of Gpc3-Specific Chimeric Antigen Receptor-Engineered Natural Killer Cells for the Treatment of Hepatocellular Carcinoma. Mol. Ther. 2018, 26, 366−378.

(646) Tsuchiya, N.; Hosono, A.; Yoshikawa, T.; Shoda, K.; Nosaka, K.; Shimomura, M.; Hara, J.; Nitani, C.; Manabe, A.; Yoshihara, H.; et al. Phase I Study of Glypican-3-Derived Peptide Vaccine Therapy for Patients with Refractory Pediatric Solid Tumors. Oncoimmunology 2018, 7, e1377872. (647) Tsuchiya, N.; Yoshikawa, T.; Fujinami, N.; Saito, K.; Mizuno, S.; Sawada, Y.; Endo, I.; Nakatsura, T. Immunological Efficacy of Glypican-3 Peptide Vaccine in Patients with Advanced Hepatocellular Carcinoma. Oncoimmunology 2017, 6, e1346764. (648) Shimizu, Y.; Suzuki, T.; Yoshikawa, T.; Tsuchiya, N.; Sawada, Y.; Endo, I.; Nakatsura, T. Cancer Immunotherapy-Targeted Glypican-3 or Neoantigens. Cancer Sci. 2018, 109, 531−541. (649) Watanabe, N.; Araki, W.; Chui, D. H.; Makifuchi, T.; Ihara, Y.; Tabira, T. Glypican-1 as an Abeta Binding Hspg in the Human Brain: Its Localization in Dig Domains and Possible Roles in the Pathogenesis of Alzheimer’s Disease. FASEB J. 2004, 18, 1013−1015. (650) O’Callaghan, P.; Sandwall, E.; Li, J. P.; Yu, H.; Ravid, R.; Guan, Z. Z.; van Kuppevelt, T. H.; Nilsson, L. N.; Ingelsson, M.; Hyman, B. T.; et al. Heparan Sulfate Accumulation with Abeta Deposits in Alzheimer’s Disease and Tg2576 Mice Is Contributed by Glial Cells. Brain Pathol. 2008, 18, 548−561. (651) Reinhard, C.; Borgers, M.; David, G.; De Strooper, B. Soluble Amyloid-Beta Precursor Protein Binds Its Cell Surface Receptor in a Cooperative Fashion with Glypican and Syndecan Proteoglycans. J. Cell Sci. 2013, 126, 4856−4861. (652) Lorentzen, A. R.; Melum, E.; Ellinghaus, E.; Smestad, C.; Mero, I. L.; Aarseth, J. H.; Myhr, K. M.; Celius, E. G.; Lie, B. A.; Karlsen, T. H.; et al. Association to the Glypican-5 Gene in Multiple Sclerosis. J. Neuroimmunol. 2010, 226, 194−197. (653) Mowry, E. M.; Carey, R. F.; Blasco, M. R.; Pelletier, J.; Duquette, P.; Villoslada, P.; Malikova, I.; Roger, E.; Kinkel, R. P.; McDonald, J.; et al. Multiple Sclerosis Susceptibility Genes: Associations with Relapse Severity and Recovery. PLoS One 2013, 8, e75416. (654) Shin, J. G.; Kim, H. J.; Park, B. L.; Bae, J. S.; Kim, L. H.; Cheong, H. S.; Shin, H. D. Putative Association of Gpc5 Polymorphism with the Risk of Inflammatory Demyelinating Diseases. J. Neurol. Sci. 2013, 335, 82−88. (655) Szatmari, T.; Otvos, R.; Hjerpe, A.; Dobra, K. Syndecan-1 in Cancer: Implications for Cell Signaling, Differentiation, and Prognostication. Dis. Markers 2015, 2015, 1−13. (656) Andersen, N. F.; Kristensen, I. B.; Preiss, B. S.; Christensen, J. H.; Abildgaard, N. Upregulation of Syndecan-1 in the Bone Marrow Microenvironment in Multiple Myeloma Is Associated with Angiogenesis. Eur. J. Haematol. 2015, 95, 211−217. (657) Szarvas, T.; Reis, H.; Kramer, G.; Shariat, S. F.; Vom Dorp, F.; Tschirdewahn, S.; Schmid, K. W.; Kovalszky, I.; Rubben, H. Enhanced Stromal Syndecan-1 Expression Is an Independent Risk Factor for Poor Survival in Bladder Cancer. Hum. Pathol. 2014, 45, 674−682. (658) Szarvas, T.; Reis, H.; Vom Dorp, F.; Tschirdewahn, S.; Niedworok, C.; Nyirady, P.; Schmid, K. W.; Rubben, H.; Kovalszky, I. Soluble Syndecan-1 (Sdc1) Serum Level as an Independent PreOperative Predictor of Cancer-Specific Survival in Prostate Cancer. Prostate 2016, 76, 977−985. (659) Yu, S.; Lv, H.; Zhang, H.; Jiang, Y.; Hong, Y.; Xia, R.; Zhang, Q.; Ju, W.; Jiang, L.; Ou, G.; et al. Heparanase-1-Induced Shedding of Heparan Sulfate from Syndecan-1 in Hepatocarcinoma Cell Facilitates Lymphatic Endothelial Cell Proliferation Via Vegf-C/Erk Pathway. Biochem. Biophys. Res. Commun. 2017, 485, 432−439. (660) Popovic, A.; Demirovic, A.; Spajic, B.; Stimac, G.; Kruslin, B.; Tomas, D. Expression and Prognostic Role of Syndecan-2 in Prostate Cancer. Prostate Cancer Prostatic Dis. 2010, 13, 78−82. (661) Ledezma, R.; Cifuentes, F.; Gallegos, I.; Fulla, J.; Ossandon, E.; Castellon, E. A.; Contreras, H. R. Altered Expression Patterns of Syndecan-1 and −2 Predict Biochemical Recurrence in Prostate Cancer. Asian J. Androl. 2011, 13, 476−480. (662) Huang, X.; Xiao, D. W.; Xu, L. Y.; Zhong, H. J.; Liao, L. D.; Xie, Z. F.; Li, E. M. Prognostic Significance of Altered Expression of BS

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Sdc2 and Cyr61 in Esophageal Squamous Cell Carcinoma. Oncol. Rep. 2009, 21, 1123−1129. (663) Baba, F.; Swartz, K.; van Buren, R.; Eickhoff, J.; Zhang, Y.; Wolberg, W.; Friedl, A. Syndecan-1 and Syndecan-4 Are Overexpressed in an Estrogen Receptor-Negative, Highly Proliferative Breast Carcinoma Subtype. Breast Cancer Res. Treat. 2006, 98, 91−98. (664) Gialeli, C.; Theocharis, A. D.; Kletsas, D.; Tzanakakis, G. N.; Karamanos, N. K. Expression of Matrix Macromolecules and Functional Properties of Egf-Responsive Colon Cancer Cells Are Inhibited by Panitumumab. Invest. New Drugs 2013, 31, 516−524. (665) Na, K. Y.; Bacchini, P.; Bertoni, F.; Kim, Y. W.; Park, Y. K. Syndecan-4 and Fibronectin in Osteosarcoma. Pathology 2012, 44, 325−330. (666) O’Connell, M. P.; Fiori, J. L.; Kershner, E. K.; Frank, B. P.; Indig, F. E.; Taub, D. D.; Hoek, K. S.; Weeraratna, A. T. Heparan Sulfate Proteoglycan Modulation of Wnt5a Signal Transduction in Metastatic Melanoma Cells. J. Biol. Chem. 2009, 284, 28704−28712. (667) Takashima, S.; Oka, Y.; Fujiki, F.; Morimoto, S.; Nakajima, H.; Nakae, Y.; Nakata, J.; Nishida, S.; Hosen, N.; Tatsumi, N.; et al. Syndecan-4 as a Biomarker to Predict Clinical Outcome for Glioblastoma Multiforme Treated with Wt1 Peptide Vaccine. Future Sci. OA 2016, 2, FSO96. (668) Labropoulou, V. T.; Skandalis, S. S.; Ravazoula, P.; Perimenis, P.; Karamanos, N. K.; Kalofonos, H. P.; Theocharis, A. D. Expression of Syndecan-4 and Correlation with Metastatic Potential in Testicular Germ Cell Tumours. BioMed Res. Int. 2013, 2013, 214864. (669) Chung, J. S.; Shiue, L. H.; Duvic, M.; Pandya, A.; Cruz, P. D., Jr.; Ariizumi, K. Sezary Syndrome Cells Overexpress Syndecan-4 Bearing Distinct Heparan Sulfate Moieties That Suppress T-Cell Activation by Binding Dc-Hil and Trapping Tgf-Beta on the Cell Surface. Blood 2011, 117, 3382−3390. (670) Orecchia, P.; Conte, R.; Balza, E.; Pietra, G.; Mingari, M. C.; Carnemolla, B. Targeting Syndecan-1, a Molecule Implicated in the Process of Vasculogenic Mimicry, Enhances the Therapeutic Efficacy of the L19-Il2 Immunocytokine in Human Melanoma Xenografts. Oncotarget 2015, 6, 37426−37442. (671) Schönfeld, K.; Herbener, P.; Zuber, C.; Häder, T.; Bernöster, K.; Uherek, C.; Schüttrumpf, J. Activity of Indatuximab Ravtansine against Triple-Negative Breast Cancer in Preclinical Tumor Models. Pharm. Res. 2018, 35, 118. (672) Schönfeld, K.; Zuber, C.; Pinkas, J.; Häder, T.; Bernöster, K.; Uherek, C. Indatuximab Ravtansine (Bt062) Combination Treatment in Multiple Myeloma: Pre-Clinical Studies. J. Hematol. Oncol. 2017, 10, 13. (673) Sondergeld, P.; van de Donk, N. W.; Richardson, P. G.; Plesner, T. Monoclonal Antibodies in Myeloma. Clin. Adv. Hematol. Oncol. 2015, 13, 599−609. (674) Vasuthasawat, A.; Yoo, E. M.; Trinh, K. R.; Lichtenstein, A.; Timmerman, J. M.; Morrison, S. L. Targeted Immunotherapy Using Anti-Cd138-Interferon Alpha Fusion Proteins and Bortezomib Results in Synergistic Protection against Multiple Myeloma. MAbs 2016, 8, 1386−1397. (675) Ikeda, H.; Hideshima, T.; Fulciniti, M.; Lutz, R. J.; Yasui, H.; Okawa, Y.; Kiziltepe, T.; Vallet, S.; Pozzi, S.; Santo, L.; et al. The Monoclonal Antibody Nbt062 Conjugated to Cytotoxic Maytansinoids Has Selective Cytotoxicity against Cd138-Positive Multiple Myeloma Cells in Vitro and in Vivo. Clin. Cancer Res. 2009, 15, 4028−4037. (676) Bae, J.; Hideshima, T.; Zhang, G. L.; Zhou, J.; Keskin, D. B.; Munshi, N. C.; Anderson, K. C. Identification and Characterization of Hla-A24-Specific Xbp1, Cd138 (Syndecan-1) and Cs1 (Slamf7) Peptides Inducing Antigens-Specific Memory Cytotoxic T Lymphocytes Targeting Multiple Myeloma. Leukemia 2018, 32, 752−764. (677) Bae, J.; Prabhala, R.; Voskertchian, A.; Brown, A.; Maguire, C.; Richardson, P.; Dranoff, G.; Anderson, K. C.; Munshi, N. C. A Multiepitope of Xbp1, Cd138 and Cs1 Peptides Induces MyelomaSpecific Cytotoxic T Lymphocytes in T Cells of Smoldering Myeloma Patients. Leukemia 2015, 29, 218−229.

(678) Jiang, H.; Zhang, W.; Shang, P.; Zhang, H.; Fu, W.; Ye, F.; Zeng, T.; Huang, H.; Zhang, X.; Sun, W.; et al. Transfection of Chimeric Anti-Cd138 Gene Enhances Natural Killer Cell Activation and Killing of Multiple Myeloma Cells. Mol. Oncol. 2014, 8, 297−310. (679) Jung, O.; Trapp-Stamborski, V.; Purushothaman, A.; Jin, H.; Wang, H.; Sanderson, R. D.; Rapraeger, A. C. Heparanase-Induced Shedding of Syndecan-1/Cd138 in Myeloma and Endothelial Cells Activates Vegfr2 and an Invasive Phenotype: Prevention by Novel Synstatins. Oncogenesis 2016, 5, e202. (680) Roemisch, J.; Gray, E.; Hoffmann, J. N.; Wiedermann, C. J. Antithrombin: A New Look at the Actions of a Serine Protease Inhibitor. Blood Coagulation Fibrinolysis 2002, 13, 657−670. (681) Teng, Y. H.; Aquino, R. S.; Park, P. W. Molecular Functions of Syndecan-1 in Disease. Matrix Biol. 2012, 31, 3−16. (682) Vanhoutte, D.; Schellings, M. W.; Gotte, M.; Swinnen, M.; Herias, V.; Wild, M. K.; Vestweber, D.; Chorianopoulos, E.; Cortes, V.; Rigotti, A.; et al. Increased Expression of Syndecan-1 Protects against Cardiac Dilatation and Dysfunction after Myocardial Infarction. Circulation 2007, 115, 475−482. (683) Wang, Z.; Li, R.; Tan, J.; Peng, L.; Wang, P.; Liu, J.; Xiong, H.; Jiang, B.; Chen, Y. Syndecan-1 Acts in Synergy with Tight Junction through Stat3 Signaling to Maintain Intestinal Mucosal Barrier and Prevent Bacterial Translocation. Inflamm. Bowel Dis. 2015, 21, 1894− 1907. (684) Yablecovitch, D.; Stein, A.; Shabat-Simon, M.; Naftali, T.; Gabay, G.; Laish, I.; Oren, A.; Konikoff, F. M. Soluble Syndecan-1 Levels Are Elevated in Patients with Inflammatory Bowel Disease. Dig. Dis. Sci. 2015, 60, 2419−2426. (685) Floer, M.; Gotte, M.; Wild, M. K.; Heidemann, J.; Gassar, E. S.; Domschke, W.; Kiesel, L.; Luegering, A.; Kucharzik, T. Enoxaparin Improves the Course of Dextran Sodium Sulfate-Induced Colitis in Syndecan-1-Deficient Mice. Am. J. Pathol. 2010, 176, 146−157. (686) Bao, X.; Moseman, E. A.; Saito, H.; Petryanik, B.; Thiriot, A.; Hatakeyama, S.; Ito, Y.; Kawashima, H.; Yamaguchi, Y.; Lowe, J. B.; et al. Endothelial Heparan Sulfate Controls Chemokine Presentation in Recruitment of Lymphocytes and Dendritic Cells to Lymph Nodes. Immunity 2010, 33, 817−829. (687) Li, Q.; Park, P. W.; Wilson, C. L.; Parks, W. C. Matrilysin Shedding of Syndecan-1 Regulates Chemokine Mobilization and Transepithelial Efflux of Neutrophils in Acute Lung Injury. Cell 2002, 111, 635−646. (688) Hayashida, K.; Parks, W. C.; Park, P. W. Syndecan-1 Shedding Facilitates the Resolution of Neutrophilic Inflammation by Removing Sequestered Cxc Chemokines. Blood 2009, 114, 3033−3043. (689) Schmidtke, M.; Karger, A.; Meerbach, A.; Egerer, R.; Stelzner, A.; Makarov, V. Binding of a N,N′-Bisheteryl Derivative of Dispirotripiperazine to Heparan Sulfate Residues on the Cell Surface Specifically Prevents Infection of Viruses from Different Families. Virology 2003, 311, 134−143. (690) Schmidtke, M.; Riabova, O.; Dahse, H. M.; Stelzner, A.; Makarov, V. Synthesis, Cytotoxicity and Antiviral Activity of N,N′Bis-5-Nitropyrimidyl Derivatives of Dispirotripiperazine. Antiviral Res. 2002, 55, 117−127. (691) Selinka, H. C.; Florin, L.; Patel, H. D.; Freitag, K.; Schmidtke, M.; Makarov, V. A.; Sapp, M. Inhibition of Transfer to Secondary Receptors by Heparan Sulfate-Binding Drug or Antibody Induces Noninfectious Uptake of Human Papillomavirus. J. Virol. 2007, 81, 10970−10980. (692) Roan, N. R.; Sowinski, S.; Munch, J.; Kirchhoff, F.; Greene, W. C. Aminoquinoline Surfen Inhibits the Action of Sevi (SemenDerived Enhancer of Viral Infection). J. Biol. Chem. 2010, 285, 1861− 1869. (693) Park, P. W.; Pier, G. B.; Hinkes, M. T.; Bernfield, M. Exploitation of Syndecan-1 Shedding by Pseudomonas Aeruginosa Enhances Virulence. Nature 2001, 411, 98−102. (694) Zhang, X.; Wu, C.; Song, J.; Gotte, M.; Sorokin, L. Syndecan1, a Cell Surface Proteoglycan, Negatively Regulates Initial Leukocyte Recruitment to the Brain across the Choroid Plexus in Murine BT

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Experimental Autoimmune Encephalomyelitis. J. Immunol. 2013, 191, 4551−4561. (695) Warford, J. R.; Lamport, A. C.; Clements, D. R.; Malone, A.; Kennedy, B. E.; Kim, Y.; Gujar, S. A.; Hoskin, D. W.; Easton, A. S. Surfen, a Proteoglycan Binding Agent, Reduces Inflammation but Inhibits Remyelination in Murine Models of Multiple Sclerosis. Acta Neuropathol. Commun. 2018, 6, 4. (696) Duckworth, C. A.; Guimond, S. E.; Sindrewicz, P.; Hughes, A. J.; French, N. S.; Lian, L. Y.; Yates, E. A.; Pritchard, D. M.; Rhodes, J. M.; Turnbull, J. E.; et al. Chemically Modified, Non-Anticoagulant Heparin Derivatives Are Potent Galectin-3 Binding Inhibitors and Inhibit Circulating Galectin-3-Promoted Metastasis. Oncotarget 2015, 6, 23671−23687. (697) Pfankuchen, D. B.; Baltes, F.; Batool, T.; Li, J. P.; Schlesinger, M.; Bendas, G. Heparin Antagonizes Cisplatin Resistance of A2780 Ovarian Cancer Cells by Affecting the Wnt Signaling Pathway. Oncotarget 2017, 8, 67553−67566. (698) Mani, K.; Belting, M.; Ellervik, U.; Falk, N.; Svensson, G.; Sandgren, S.; Cheng, F.; Fransson, L. A. Tumor Attenuation by 2(6Hydroxynaphthyl)-Beta-D-Xylopyranoside Requires Priming of Heparan Sulfate and Nuclear Targeting of the Products. Glycobiology 2004, 14, 387−397. (699) Brule, S.; Friand, V.; Sutton, A.; Baleux, F.; Gattegno, L.; Charnaux, N. Glycosaminoglycans and Syndecan-4 Are Involved in Sdf-1/Cxcl12-Mediated Invasion of Human Epitheloid Carcinoma Hela Cells. Biochim. Biophys. Acta, Gen. Subj. 2009, 1790, 1643−1650. (700) Raman, K.; Kuberan, B. Click-Xylosides Mitigate Glioma Cell Invasion in Vitro. Mol. BioSyst. 2010, 6, 1800−1802. (701) Purushothaman, A.; Toole, B. P. Serglycin Proteoglycan Is Required for Multiple Myeloma Cell Adhesion, in Vivo Growth, and Vascularization. J. Biol. Chem. 2014, 289, 5499−5509. (702) Theocharis, A. D.; Seidel, C.; Borset, M.; Dobra, K.; Baykov, V.; Labropoulou, V.; Kanakis, I.; Dalas, E.; Karamanos, N. K.; Sundan, A.; et al. Serglycin Constitutively Secreted by Myeloma Plasma Cells Is a Potent Inhibitor of Bone Mineralization in Vitro. J. Biol. Chem. 2006, 281, 35116−35128. (703) Niemann, C. U.; Kjeldsen, L.; Ralfkiaer, E.; Jensen, M. K.; Borregaard, N. Serglycin Proteoglycan in Hematologic Malignancies: A Marker of Acute Myeloid Leukemia. Leukemia 2007, 21, 2406− 2410. (704) Korpetinou, A.; Papachristou, D. J.; Lampropoulou, A.; Bouris, P.; Labropoulou, V. T.; Noulas, A.; Karamanos, N. K.; Theocharis, A. D. Increased Expression of Serglycin in Specific Carcinomas and Aggressive Cancer Cell Lines. BioMed Res. Int. 2015, 2015, 690721. (705) Korpetinou, A.; Skandalis, S. S.; Moustakas, A.; Happonen, K. E.; Tveit, H.; Prydz, K.; Labropoulou, V. T.; Giannopoulou, E.; Kalofonos, H. P.; Blom, A. M.; et al. Serglycin Is Implicated in the Promotion of Aggressive Phenotype of Breast Cancer Cells. PLoS One 2013, 8, e78157. (706) Zhang, Z.; Deng, Y.; Zheng, G.; Jia, X.; Xiong, Y.; Luo, K.; Qiu, Q.; Qiu, N.; Yin, J.; Lu, M.; et al. Srgn-Tgfbeta2 Regulatory Loop Confers Invasion and Metastasis in Triple-Negative Breast Cancer. Oncogenesis 2017, 6, e360. (707) Guo, J. Y.; Hsu, H. S.; Tyan, S. W.; Li, F. Y.; Shew, J. Y.; Lee, W. H.; Chen, J. Y. Serglycin in Tumor Microenvironment Promotes Non-Small Cell Lung Cancer Aggressiveness in a Cd44-Dependent Manner. Oncogene 2017, 36, 2457−2471. (708) He, L.; Zhou, X.; Qu, C.; Tang, Y.; Zhang, Q.; Hong, J. Serglycin (Srgn) Overexpression Predicts Poor Prognosis in Hepatocellular Carcinoma Patients. Med. Oncol. 2013, 30, 707. (709) Li, X. J.; Ong, C. K.; Cao, Y.; Xiang, Y. Q.; Shao, J. Y.; Ooi, A.; Peng, L. X.; Lu, W. H.; Zhang, Z.; Petillo, D.; et al. Serglycin Is a Theranostic Target in Nasopharyngeal Carcinoma That Promotes Metastasis. Cancer Res. 2011, 71, 3162−3172. (710) Lorenzi, L.; Doring, C.; Rausch, T.; Benes, V.; Lonardi, S.; Bugatti, M.; Campo, E.; Cabecadas, J.; Simonitsch-Klupp, I.; Borges, A.; et al. Identification of Novel Follicular Dendritic Cell Sarcoma

Markers, Fdcsp and Srgn, by Whole Transcriptome Sequencing. Oncotarget 2017, 8, 16463−16472. (711) Roy, A.; Attarha, S.; Weishaupt, H.; Edqvist, P. H.; Swartling, F. J.; Bergqvist, M.; Siebzehnrubl, F. A.; Smits, A.; Ponten, F.; Tchougounova, E. Serglycin as a Potential Biomarker for Glioma: Association of Serglycin Expression, Extent of Mast Cell Recruitment and Glioblastoma Progression. Oncotarget 2017, 8, 24815−24827. (712) Skliris, A.; Happonen, K. E.; Terpos, E.; Labropoulou, V.; Borset, M.; Heinegard, D.; Blom, A. M.; Theocharis, A. D. Serglycin Inhibits the Classical and Lectin Pathways of Complement Via Its Glycosaminoglycan Chains: Implications for Multiple Myeloma. Eur. J. Immunol. 2011, 41, 437−449. (713) Skliris, A.; Labropoulou, V. T.; Papachristou, D. J.; Aletras, A.; Karamanos, N. K.; Theocharis, A. D. Cell-Surface Serglycin Promotes Adhesion of Myeloma Cells to Collagen Type I and Affects the Expression of Matrix Metalloproteinases. FEBS J. 2013, 280, 2342− 2352. (714) Clausen, T. M.; Pereira, M. A.; Al Nakouzi, N.; Oo, H. Z.; Agerbaek, M. O.; Lee, S.; Orum-Madsen, M. S.; Kristensen, A. R.; ElNaggar, A.; Grandgenett, P. M.; et al. Oncofetal Chondroitin Sulfate Glycosaminoglycans Are Key Players in Integrin Signaling and Tumor Cell Motility. Mol. Cancer Res. 2016, 14, 1288−1299. (715) Bouris, P.; Sopaki-Valalaki, N.; Kolokotroni, A.; Moustakas, A.; Kapoor, A.; Iozzo, R. V.; Karamanos, N. K.; Theocharis, A. D.; Manou, D. Serglycin Promotes Breast Cancer Cell Aggressiveness: Induction of Epithelial to Mesenchymal Transition, Proteolytic Activity and Il-8 Signaling. Matrix Biol. 2018, DOI: 10.1016/ j.matbio.2018.05.011. (716) Chu, Q.; Huang, H.; Huang, T.; Cao, L.; Peng, L.; Shi, S.; Zheng, L.; Xu, L.; Zhang, S.; Huang, J.; et al. Extracellular Serglycin Upregulates the Cd44 Receptor in an Autocrine Manner to Maintain Self-Renewal in Nasopharyngeal Carcinoma Cells by Reciprocally Activating the Mapk/Beta-Catenin Axis. Cell Death Dis. 2016, 7, e2456. (717) Hamilton, A.; Basic, V.; Andersson, S.; Abrink, M.; Ringvall, M. Loss of Serglycin Promotes Primary Tumor Growth and Vessel Functionality in the Rip1-Tag2Mouse Model for Spontaneous Insulinoma Formation. PLoS One 2015, 10, e0126688. (718) Roy, A.; Femel, J.; Huijbers, E. J.; Spillmann, D.; Larsson, E.; Ringvall, M.; Olsson, A. K.; Abrink, M. Targeting Serglycin Prevents Metastasis in Murine Mammary Carcinoma. PLoS One 2016, 11, e0156151. (719) Purushothaman, A.; Bandari, S. K.; Chandrashekar, D. S.; Jones, R. J.; Lee, H. C.; Weber, D. M.; Orlowski, R. Z. Chondroitin Sulfate Proteoglycan Serglycin Influences Protein Cargo Loading and Functions of Tumor-Derived Exosomes. Oncotarget 2017, 8, 73723− 73732. (720) Kolset, S. O.; Pejler, G. Serglycin: A Structural and Functional Chameleon with Wide Impact on Immune Cells. J. Immunol. 2011, 187, 4927−4933. (721) Sutton, V. R.; Brennan, A. J.; Ellis, S.; Danne, J.; Thia, K.; Jenkins, M. R.; Voskoboinik, I.; Pejler, G.; Johnstone, R. W.; Andrews, D. M.; et al. Serglycin Determines Secretory Granule Repertoire and Regulates Natural Killer Cell and Cytotoxic T Lymphocyte Cytotoxicity. FEBS J. 2016, 283, 947−961. (722) Abrink, M.; Grujic, M.; Pejler, G. Serglycin Is Essential for Maturation of Mast Cell Secretory Granule. J. Biol. Chem. 2004, 279, 40897−40905. (723) Roy, A.; Sawesi, O.; Pettersson, U.; Dagalv, A.; Kjellen, L.; Lunden, A.; Abrink, M. Serglycin Proteoglycans Limit Enteropathy in Trichinella Spiralis-Infected Mice. BMC Immunol. 2016, 17, 15. (724) D’Ascola, A.; Scuruchi, M.; Avenoso, A.; Bruschetta, G.; Campo, S.; Mandraffino, G.; Campo, G. M. Serglycin Is Involved in Inflammatory Response in Articular Mouse Chondrocytes. Biochem. Biophys. Res. Commun. 2018, 499, 506−512. (725) Yanez-Mo, M.; Siljander, P. R.; Andreu, Z.; Zavec, A. B.; Borras, F. E.; Buzas, E. I.; Buzas, K.; Casal, E.; Cappello, F.; Carvalho, J.; et al. Biological Properties of Extracellular Vesicles and Their Physiological Functions. J. Extracell. Vesicles 2015, 4, 27066. BU

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Cleavages of a Linker Segment. J. Biol. Chem. 2008, 283, 18167− 18176. (746) Vlodavsky, I.; Singh, P.; Boyango, I.; Gutter-Kapon, L.; Elkin, M.; Sanderson, R. D.; Ilan, N. Heparanase: From Basic Research to Therapeutic Applications in Cancer and Inflammation. Drug Resist. Updates 2016, 29, 54−75. (747) Gingis-Velitski, S.; Zetser, A.; Flugelman, M. Y.; Vlodavsky, I.; Ilan, N. Heparanase Induces Endothelial Cell Migration Via Protein Kinase B/Akt Activation. J. Biol. Chem. 2004, 279, 23536−23541. (748) Purushothaman, A.; Chen, L.; Yang, Y.; Sanderson, R. D. Heparanase Stimulation of Protease Expression Implicates It as a Master Regulator of the Aggressive Tumor Phenotype in Myeloma. J. Biol. Chem. 2008, 283, 32628−32636. (749) Purushothaman, A.; Uyama, T.; Kobayashi, F.; Yamada, S.; Sugahara, K.; Rapraeger, A. C.; Sanderson, R. D. HeparanaseEnhanced Shedding of Syndecan-1 by Myeloma Cells Promotes Endothelial Invasion and Angiogenesis. Blood 2010, 115, 2449−2457. (750) Cohen-Kaplan, V.; Naroditsky, I.; Zetser, A.; Ilan, N.; Vlodavsky, I.; Doweck, I. Heparanase Induces Vegf C and Facilitates Tumor Lymphangiogenesis. Int. J. Cancer 2008, 123, 2566−2573. (751) Ramani, V. C.; Yang, Y.; Ren, Y.; Nan, L.; Sanderson, R. D. Heparanase Plays a Dual Role in Driving Hepatocyte Growth Factor (Hgf) Signaling by Enhancing Hgf Expression and Activity. J. Biol. Chem. 2011, 286, 6490−6499. (752) Purushothaman, A.; Bandari, S. K.; Liu, J.; Mobley, J. A.; Brown, E. E.; Sanderson, R. D. Fibronectin on the Surface of Myeloma Cell-Derived Exosomes Mediates Exosome-Cell Interactions. J. Biol. Chem. 2016, 291, 1652−1663. (753) Bandari, S. K.; Purushothaman, A.; Ramani, V. C.; Brinkley, G. J.; Chandrashekar, D. S.; Varambally, S.; Mobley, J. A.; Zhang, Y.; Brown, E. E.; Vlodavsky, I.; et al. Chemotherapy Induces Secretion of Exosomes Loaded with Heparanase That Degrades Extracellular Matrix and Impacts Tumor and Host Cell Behavior. Matrix Biol. 2018, 65, 104−118. (754) Liao, J.; Kapadia, V. S.; Brown, L. S.; Cheong, N.; Longoria, C.; Mija, D.; Ramgopal, M.; Mirpuri, J.; McCurnin, D. C.; Savani, R. C. The Nlrp3 Inflammasome Is Critically Involved in the Development of Bronchopulmonary Dysplasia. Nat. Commun. 2015, 6, 8977. (755) Csoka, A. B.; Stern, R. Hypotheses on the Evolution of Hyaluronan: A Highly Ironic Acid. Glycobiology 2013, 23, 398−411. (756) Wessels, M. R.; Moses, A. E.; Goldberg, J. B.; DiCesare, T. J. Hyaluronic Acid Capsule Is a Virulence Factor for Mucoid Group a Streptococci. Proc. Natl. Acad. Sci. U. S. A. 1991, 88, 8317−8321. (757) Stern, R. Hyaluronan Catabolism: A New Metabolic Pathway. Eur. J. Cell Biol. 2004, 83, 317−325. (758) Bohaumilitzky, L.; Huber, A. K.; Stork, E. M.; Wengert, S.; Woelfl, F.; Boehm, H. A Trickster in Disguise: Hyaluronan’s Ambivalent Roles in the Matrix. Front. Oncol. 2017, 7, 242. (759) Iijima, J.; Konno, K.; Itano, N. Inflammatory Alterations of the Extracellular Matrix in the Tumor Microenvironment. Cancers 2011, 3, 3189−3205. (760) Jiang, D.; Liang, J.; Fan, J.; Yu, S.; Chen, S.; Luo, Y.; Prestwich, G. D.; Mascarenhas, M. M.; Garg, H. G.; Quinn, D. A.; et al. Regulation of Lung Injury and Repair by Toll-Like Receptors and Hyaluronan. Nat. Med. 2005, 11, 1173−1179. (761) Jiang, D.; Liang, J.; Noble, P. W. Hyaluronan as an Immune Regulator in Human Diseases. Physiol. Rev. 2011, 91, 221−264. (762) Monslow, J.; Govindaraju, P.; Puré, E. Hyaluronan - a Functional and Structural Sweet Spot in the Tissue Microenvironment. Front. Immunol. 2015, 6, 231. (763) Rugg, M. S.; Willis, A. C.; Mukhopadhyay, D.; Hascall, V. C.; Fries, E.; Fülöp, C.; Milner, C. M.; Day, A. J. Characterization of Complexes Formed between Tsg-6 and Inter-Alpha-Inhibitor That Act as Intermediates in the Covalent Transfer of Heavy Chains onto Hyaluronan. J. Biol. Chem. 2005, 280, 25674−25686. (764) de la Motte, C. A.; Hascall, V. C.; Drazba, J.; Bandyopadhyay, S. K.; Strong, S. A. Mononuclear Leukocytes Bind to Specific Hyaluronan Structures on Colon Mucosal Smooth Muscle Cells Treated with Polyinosinic Acid:Polycytidylic Acid: Inter-Alpha-

(726) Shang, M.; Ji, J. S.; Song, C.; Gao, B. J.; Jin, J. G.; Kuo, W. P.; Kang, H. Extracellular Vesicles: A Brief Overview and Its Role in Precision Medicine. Methods Mol. Biol. 2017, 1660, 1−14. (727) Colombo, M.; Raposo, G.; Thery, C. Biogenesis, Secretion, and Intercellular Interactions of Exosomes and Other Extracellular Vesicles. Annu. Rev. Cell Dev. Biol. 2014, 30, 255−289. (728) Witwer, K. W.; Buzas, E. I.; Bemis, L. T.; Bora, A.; Lasser, C.; Lotvall, J.; Nolte-’t Hoen, E. N.; Piper, M. G.; Sivaraman, S.; Skog, J.et al. Standardization of Sample Collection, Isolation and Analysis Methods in Extracellular Vesicle Research. J. Extracell. Vesicles 2013, 2, 20360. (729) Stoorvogel, W.; Strous, G. J.; Geuze, H. J.; Oorschot, V.; Schwartz, A. L. Late Endosomes Derive from Early Endosomes by Maturation. Cell 1991, 65, 417−427. (730) Raposo, G.; Nijman, H. W.; Stoorvogel, W.; Liejendekker, R.; Harding, C. V.; Melief, C. J.; Geuze, H. J. B Lymphocytes Secrete Antigen-Presenting Vesicles. J. Exp. Med. 1996, 183, 1161−1172. (731) Raposo, G.; Stoorvogel, W. Extracellular Vesicles: Exosomes, Microvesicles, and Friends. J. Cell Biol. 2013, 200, 373−383. (732) Mobius, W.; van Donselaar, E.; Ohno-Iwashita, Y.; Shimada, Y.; Heijnen, H. F.; Slot, J. W.; Geuze, H. J. Recycling Compartments and the Internal Vesicles of Multivesicular Bodies Harbor Most of the Cholesterol Found in the Endocytic Pathway. Traffic 2003, 4, 222− 231. (733) Hanson, P. I.; Cashikar, A. Multivesicular Body Morphogenesis. Annu. Rev. Cell Dev. Biol. 2012, 28, 337−362. (734) Baietti, M. F.; Zhang, Z.; Mortier, E.; Melchior, A.; Degeest, G.; Geeraerts, A.; Ivarsson, Y.; Depoortere, F.; Coomans, C.; Vermeiren, E.; et al. Syndecan-Syntenin-Alix Regulates the Biogenesis of Exosomes. Nat. Cell Biol. 2012, 14, 677−685. (735) Grootjans, J. J.; Reekmans, G.; Ceulemans, H.; David, G. Syntenin-Syndecan Binding Requires Syndecan-Synteny and the CoOperation of Both Pdz Domains of Syntenin. J. Biol. Chem. 2000, 275, 19933−19941. (736) Mathivanan, S.; Simpson, R. J. Exocarta: A Compendium of Exosomal Proteins and Rna. Proteomics 2009, 9, 4997−5000. (737) Imjeti, N. S.; Menck, K.; Egea-Jimenez, A. L.; Lecointre, C.; Lembo, F.; Bouguenina, H.; Badache, A.; Ghossoub, R.; David, G.; Roche, S.; et al. Syntenin Mediates Src Function in Exosomal Cell-toCell Communication. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 12495−12500. (738) Roucourt, B.; Meeussen, S.; Bao, J.; Zimmermann, P.; David, G. Heparanase Activates the Syndecan-Syntenin-Alix Exosome Pathway. Cell Res. 2015, 25, 412−428. (739) Migliorini, E.; Thakar, D.; Kuhnle, J.; Sadir, R.; Dyer, D. P.; Li, Y.; Sun, C.; Volkman, B. F.; Handel, T. M.; Coche-Guerente, L.et al. Cytokines and Growth Factors Cross-Link Heparan Sulfate. Open Biol. 2015, 5, 150046. (740) Fang, Y.; Wu, N.; Gan, X.; Yan, W.; Morrell, J. C.; Gould, S. J. Higher-Order Oligomerization Targets Plasma Membrane Proteins and Hiv Gag to Exosomes. PLoS Biol. 2007, 5, e158. (741) Thompson, C. A.; Purushothaman, A.; Ramani, V. C.; Vlodavsky, I.; Sanderson, R. D. Heparanase Regulates Secretion, Composition, and Function of Tumor Cell-Derived Exosomes. J. Biol. Chem. 2013, 288, 10093−10099. (742) Fux, L.; Ilan, N.; Sanderson, R. D.; Vlodavsky, I. Heparanase: Busy at the Cell Surface. Trends Biochem. Sci. 2009, 34, 511−519. (743) Zetser, A.; Levy-Adam, F.; Kaplan, V.; Gingis-Velitski, S.; Bashenko, Y.; Schubert, S.; Flugelman, M. Y.; Vlodavsky, I.; Ilan, N. Processing and Activation of Latent Heparanase Occurs in Lysosomes. J. Cell Sci. 2004, 117, 2249−2258. (744) Vreys, V.; Delande, N.; Zhang, Z.; Coomans, C.; Roebroek, A.; Durr, J.; David, G. Cellular Uptake of Mammalian Heparanase Precursor Involves Low Density Lipoprotein Receptor-Related Proteins, Mannose 6-Phosphate Receptors, and Heparan Sulfate Proteoglycans. J. Biol. Chem. 2005, 280, 33141−33148. (745) Abboud-Jarrous, G.; Atzmon, R.; Peretz, T.; Palermo, C.; Gadea, B. B.; Joyce, J. A.; Vlodavsky, I. Cathepsin L Is Responsible for Processing and Activation of Proheparanase through Multiple BV

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Trypsin Inhibitor Is Crucial to Structure and Function. Am. J. Pathol. 2003, 163, 121−133. (765) Petrey, A. C.; de la Motte, C. A. Hyaluronan, a Crucial Regulator of Inflammation. Front. Immunol. 2014, 5, 101. (766) Zhuo, L.; Yoneda, M.; Zhao, M.; Yingsung, W.; Yoshida, N.; Kitagawa, Y.; Kawamura, K.; Suzuki, T.; Kimata, K. Defect in ShapHyaluronan Complex Causes Severe Female Infertility. A Study by Inactivation of the Bikunin Gene in Mice. J. Biol. Chem. 2001, 276, 7693−7696. (767) Sato, H.; Kajikawa, S.; Kuroda, S.; Horisawa, Y.; Nakamura, N.; Kaga, N.; Kakinuma, C.; Kato, K.; Morishita, H.; Niwa, H.; et al. Impaired Fertility in Female Mice Lacking Urinary Trypsin Inhibitor. Biochem. Biophys. Res. Commun. 2001, 281, 1154−1160. (768) Jessen, T. E.; Ødum, L. Role of Tumour Necrosis Factor Stimulated Gene 6 (Tsg-6) in the Coupling of Inter-Alpha-Trypsin Inhibitor to Hyaluronan in Human Follicular Fluid. Reproduction 2003, 125, 27−31. (769) Weigel, P. H.; DeAngelis, P. L. Hyaluronan Synthases: A Decade-Plus of Novel Glycosyltransferases. J. Biol. Chem. 2007, 282, 36777−36781. (770) Weigel, P. H.; Baggenstoss, B. A.; Washburn, J. L. Hyaluronan Synthase Assembles Hyaluronan on a [Glcnac(B1,4)]N-Glcnac(A1→)Udp Primer and Hyaluronan Retains This Residual Chitin Oligomer as a Cap at the Nonreducing End. Glycobiology 2017, 27, 536−554. (771) Lee, J. Y.; Spicer, A. P. Hyaluronan: A Multifunctional, Megadalton, Stealth Molecule. Curr. Opin. Cell Biol. 2000, 12, 581− 586. (772) Fraser, J. R.; Laurent, T. C.; Laurent, U. B. Hyaluronan: Its Nature, Distribution, Functions and Turnover. J. Intern. Med. 1997, 242, 27−33. (773) Yoshida, M.; Itano, N.; Yamada, Y.; Kimata, K. In Vitro Synthesis of Hyaluronan by a Single Protein Derived from Mouse Has1 Gene and Characterization of Amino Acid Residues Essential for the Activity. J. Biol. Chem. 2000, 275, 497−506. (774) Blackburn, M. R.; Hubbard, C.; Kiessling, V.; Bi, Y.; Kloss, B.; Tamm, L. K.; Zimmer, J. Distinct Reaction Mechanisms for Hyaluronan Biosynthesis in Different Kingdoms of Life. Glycobiology 2018, 28, 108−121. (775) Bi, Y.; Hubbard, C.; Purushotham, P.; Zimmer, J. Insights into the Structure and Function of Membrane-Integrated Processive Glycosyltransferases. Curr. Opin. Struct. Biol. 2015, 34, 78−86. (776) Itano, N.; Kimata, K. Mammalian Hyaluronan Synthases. IUBMB Life 2002, 54, 195−199. (777) Suzuki, M.; Asplund, T.; Yamashita, H.; Heldin, C. H.; Heldin, P. Stimulation of Hyaluronan Biosynthesis by Platelet-Derived Growth Factor-Bb and Transforming Growth Factor-Beta 1 Involves Activation of Protein Kinase C. Biochem. J. 1995, 307 (3), 817−821. (778) Vigetti, D.; Clerici, M.; Deleonibus, S.; Karousou, E.; Viola, M.; Moretto, P.; Heldin, P.; Hascall, V. C.; De Luca, G.; Passi, A. Hyaluronan Synthesis Is Inhibited by Adenosine MonophosphateActivated Protein Kinase through the Regulation of Has2 Activity in Human Aortic Smooth Muscle Cells. J. Biol. Chem. 2011, 286, 7917− 7924. (779) Vigetti, D.; Deleonibus, S.; Moretto, P.; Karousou, E.; Viola, M.; Bartolini, B.; Hascall, V. C.; Tammi, M.; De Luca, G.; Passi, A. Role of Udp-N-Acetylglucosamine (Glcnac) and O-Glcnacylation of Hyaluronan Synthase 2 in the Control of Chondroitin Sulfate and Hyaluronan Synthesis. J. Biol. Chem. 2012, 287, 35544−35555. (780) Karousou, E.; Kamiryo, M.; Skandalis, S. S.; Ruusala, A.; Asteriou, T.; Passi, A.; Yamashita, H.; Hellman, U.; Heldin, C. H.; Heldin, P. The Activity of Hyaluronan Synthase 2 Is Regulated by Dimerization and Ubiquitination. J. Biol. Chem. 2010, 285, 23647− 23654. (781) Deen, A. J.; Rilla, K.; Oikari, S.; Kärnä, R.; Bart, G.; Häyrinen, J.; Bathina, A. R.; Ropponen, A.; Makkonen, K.; Tammi, R. H.; et al. Rab10-Mediated Endocytosis of the Hyaluronan Synthase Has3 Regulates Hyaluronan Synthesis and Cell Adhesion to Collagen. J. Biol. Chem. 2014, 289, 8375−8389.

(782) Laurent, T. C.; Fraser, J. R. Hyaluronan. FASEB J. 1992, 6, 2397−2404. (783) Vigetti, D.; Moretto, P.; Viola, M.; Genasetti, A.; Rizzi, M.; Karousou, E.; Pallotti, F.; De Luca, G.; Passi, A. Matrix Metalloproteinase 2 and Tissue Inhibitors of Metalloproteinases Regulate Human Aortic Smooth Muscle Cell Migration During in Vitro Aging. FASEB J. 2006, 20, 1118−1130. (784) Motolese, A.; Vignati, F.; Brambilla, R.; Cerati, M.; Passi, A. Interaction between a Regenerative Matrix and Wound Bed in Nonhealing Ulcers: Results with 16 Cases. BioMed Res. Int. 2013, 2013, 1, . (785) Vigetti, D.; Genasetti, A.; Karousou, E.; Viola, M.; Clerici, M.; Bartolini, B.; Moretto, P.; De Luca, G.; Hascall, V. C.; Passi, A. Modulation of Hyaluronan Synthase Activity in Cellular Membrane Fractions. J. Biol. Chem. 2009, 284, 30684−30694. (786) Toole, B. P. Hyaluronan: From Extracellular Glue to Pericellular Cue. Nat. Rev. Cancer 2004, 4, 528−539. (787) Vigetti, D.; Karousou, E.; Viola, M.; Deleonibus, S.; De Luca, G.; Passi, A. Hyaluronan: Biosynthesis and Signaling. Biochim. Biophys. Acta, Gen. Subj. 2014, 1840, 2452−2459. (788) Merrilees, M. J.; Beaumont, B. W.; Braun, K. R.; Thomas, A. C.; Kang, I.; Hinek, A.; Passi, A.; Wight, T. N. Neointima Formed by Arterial Smooth Muscle Cells Expressing Versican Variant V3 Is Resistant to Lipid and Macrophage Accumulation. Arterioscler., Thromb., Vasc. Biol. 2011, 31, 1309−1316. (789) Bollyky, P. L.; Bogdani, M.; Bollyky, J. B.; Hull, R. L.; Wight, T. N. The Role of Hyaluronan and the Extracellular Matrix in Islet Inflammation and Immune Regulation. Curr. Diabetes Rep. 2012, 12, 471−480. (790) Nagy, N.; de la Zerda, A.; Kaber, G.; Johnson, P. Y.; Hu, K. H.; Kratochvil, M. J.; Yadava, K.; Zhao, W.; Cui, Y.; Navarro, G.; et al. Hyaluronan Content Governs Tissue Stiffness in Pancreatic Islet Inflammation. J. Biol. Chem. 2018, 293, 567−578. (791) Spicer, A. P.; McDonald, J. A. Characterization and Molecular Evolution of a Vertebrate Hyaluronan Synthase Gene Family. J. Biol. Chem. 1998, 273, 1923−1932. (792) Pienimaki, J. P.; Rilla, K.; Fulop, C.; Sironen, R. K.; Karvinen, S.; Pasonen, S.; Lammi, M. J.; Tammi, R.; Hascall, V. C.; Tammi, M. I. Epidermal Growth Factor Activates Hyaluronan Synthase 2 in Epidermal Keratinocytes and Increases Pericellular and Intracellular Hyaluronan. J. Biol. Chem. 2001, 276, 20428−20435. (793) Pasonen-Seppänen, S.; Karvinen, S.; Törrönen, K.; Hyttinen, J. M.; Jokela, T.; Lammi, M. J.; Tammi, M. I.; Tammi, R. Egf Upregulates, Whereas Tgf-Beta Downregulates, the Hyaluronan Synthases Has2 and Has3 in Organotypic Keratinocyte Cultures: Correlations with Epidermal Proliferation and Differentiation. J. Invest. Dermatol. 2003, 120, 1038−1044. (794) Karvinen, S.; Pasonen-Seppänen, S.; Hyttinen, J. M.; Pienimäki, J. P.; Törrönen, K.; Jokela, T. A.; Tammi, M. I.; Tammi, R. Keratinocyte Growth Factor Stimulates Migration and Hyaluronan Synthesis in the Epidermis by Activation of Keratinocyte Hyaluronan Synthases 2 and 3. J. Biol. Chem. 2003, 278, 49495−49504. (795) Jacobson, A.; Brinck, J.; Briskin, M. J.; Spicer, A. P.; Heldin, P. Expression of Human Hyaluronan Synthases in Response to External Stimuli. Biochem. J. 2000, 348 (1), 29−35. (796) Saavalainen, K.; Pasonen-Seppänen, S.; Dunlop, T. W.; Tammi, R.; Tammi, M. I.; Carlberg, C. The Human Hyaluronan Synthase 2 Gene Is a Primary Retinoic Acid and Epidermal Growth Factor Responding Gene. J. Biol. Chem. 2005, 280, 14636−14644. (797) Wang, H. S.; Tung, W. H.; Tang, K. T.; Wong, Y. K.; Huang, G. J.; Wu, J. C.; Guo, Y. J.; Chen, C. C. Tgf-Beta Induced Hyaluronan Synthesis in Orbital Fibroblasts Involves Protein Kinase C Betaii Activation in Vitro. J. Cell. Biochem. 2005, 95, 256−267. (798) Monslow, J.; Sato, N.; Mack, J. A.; Maytin, E. V. WoundingInduced Synthesis of Hyaluronic Acid in Organotypic Epidermal Cultures Requires the Release of Heparin-Binding Egf and Activation of the Egfr. J. Invest. Dermatol. 2009, 129, 2046−2058. BW

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(799) Yarema, K. J.; Bertozzi, C. R. Characterizing Glycosylation Pathways. Genome Biol. 2001, 2, DOI: 10.1186/gb-2001-2-5reviews0004. (800) Hadley, B.; Maggioni, A.; Ashikov, A.; Day, C. J.; Haselhorst, T.; Tiralongo, J. Structure and Function of Nucleotide Sugar Transporters: Current Progress. Comput. Struct. Biotechnol. J. 2014, 10, 23−32. (801) Gerardy-Schahn, R.; Oelmann, S.; Bakker, H. Nucleotide Sugar Transporters: Biological and Functional Aspects. Biochimie 2001, 83, 775−782. (802) Marshall, S.; Nadeau, O.; Yamasaki, K. Dynamic Actions of Glucose and Glucosamine on Hexosamine Biosynthesis in Isolated Adipocytes: Differential Effects on Glucosamine 6-Phosphate, Udp-NAcetylglucosamine, and Atp Levels. J. Biol. Chem. 2004, 279, 35313− 35319. (803) Piccioni, F.; Malvicini, M.; Garcia, M. G.; Rodriguez, A.; Atorrasagasti, C.; Kippes, N.; Piedra Buena, I. T.; Rizzo, M. M.; Bayo, J.; Aquino, J.; et al. Antitumor Effects of Hyaluronic Acid Inhibitor 4Methylumbelliferone in an Orthotopic Hepatocellular Carcinoma Model in Mice. Glycobiology 2012, 22, 400−410. (804) Vigetti, D.; Rizzi, M.; Moretto, P.; Deleonibus, S.; Dreyfuss, J. M.; Karousou, E.; Viola, M.; Clerici, M.; Hascall, V. C.; Ramoni, M. F.; et al. Glycosaminoglycans and Glucose Prevent Apoptosis in 4Methylumbelliferone-Treated Human Aortic Smooth Muscle Cells. J. Biol. Chem. 2011, 286, 34497−34503. (805) Torres, C. R.; Hart, G. W. Topography and Polypeptide Distribution of Terminal N-Acetylglucosamine Residues on the Surfaces of Intact Lymphocytes. Evidence for O-Linked Glcnac. J. Biol. Chem. 1984, 259, 3308−3317. (806) Dias, W. B.; Hart, G. W. O-Glcnac Modification in Diabetes and Alzheimer’s Disease. Mol. BioSyst. 2007, 3, 766−772. (807) Hart, G. W.; Housley, M. P.; Slawson, C. Cycling of O-Linked Beta-N-Acetylglucosamine on Nucleocytoplasmic Proteins. Nature 2007, 446, 1017−1022. (808) Lewis, B. A.; Hanover, J. A. O-Glcnac and the Epigenetic Regulation of Gene Expression. J. Biol. Chem. 2014, 289, 34440− 34448. (809) Vigetti, D.; Viola, M.; Karousou, E.; De Luca, G.; Passi, A. Metabolic Control of Hyaluronan Synthases. Matrix Biol. 2014, 35, 8−13. (810) Siiskonen, H.; Oikari, S.; Pasonen-Seppänen, S.; Rilla, K. Hyaluronan Synthase 1: A Mysterious Enzyme with Unexpected Functions. Front. Immunol. 2015, 6, 43. (811) Kultti, A.; Rilla, K.; Tiihonen, R.; Spicer, A. P.; Tammi, R. H.; Tammi, M. I. Hyaluronan Synthesis Induces Microvillus-Like Cell Surface Protrusions. J. Biol. Chem. 2006, 281, 15821−15828. (812) Sainio, A.; Jokela, T.; Tammi, M. I.; Järveläinen, H. Hyperglycemic Conditions Modulate Connective Tissue Reorganization by Human Vascular Smooth Muscle Cells through Stimulation of Hyaluronan Synthesis. Glycobiology 2010, 20, 1117−1126. (813) Tammi, R. H.; Passi, A. G.; Rilla, K.; Karousou, E.; Vigetti, D.; Makkonen, K.; Tammi, M. I. Transcriptional and Post-Translational Regulation of Hyaluronan Synthesis. FEBS J. 2011, 278, 1419−1428. (814) Majors, A. K.; Austin, R. C.; de la Motte, C. A.; Pyeritz, R. E.; Hascall, V. C.; Kessler, S. P.; Sen, G.; Strong, S. A. Endoplasmic Reticulum Stress Induces Hyaluronan Deposition and Leukocyte Adhesion. J. Biol. Chem. 2003, 278, 47223−47231. (815) Viola, M.; Bartolini, B.; Vigetti, D.; Karousou, E.; Moretto, P.; Deleonibus, S.; Sawamura, T.; Wight, T. N.; Hascall, V. C.; De Luca, G.; et al. Oxidized Low Density Lipoprotein (Ldl) Affects Hyaluronan Synthesis in Human Aortic Smooth Muscle Cells. J. Biol. Chem. 2013, 288, 29595−29603. (816) Tlapak-Simmons, V. L.; Baron, C. A.; Weigel, P. H. Characterization of the Purified Hyaluronan Synthase from Streptococcus Equisimilis. Biochemistry 2004, 43, 9234−9242. (817) Ontong, P.; Hatada, Y.; Taniguchi, S.; Kakizaki, I.; Itano, N. Effect of a Cholesterol-Rich Lipid Environment on the Enzymatic Activity of Reconstituted Hyaluronan Synthase. Biochem. Biophys. Res. Commun. 2014, 443, 666−671.

(818) DeAngelis, P. L. Monodisperse Hyaluronan Polymers: Synthesis and Potential Applications. Curr. Pharm. Biotechnol. 2008, 9, 246−248. (819) Jing, W.; DeAngelis, P. L. Synchronized Chemoenzymatic Synthesis of Monodisperse Hyaluronan Polymers. J. Biol. Chem. 2004, 279, 42345−42349. (820) Baggenstoss, B. A.; Harris, E. N.; Washburn, J. L.; Medina, A. P.; Nguyen, L.; Weigel, P. H. Hyaluronan Synthase Control of Synthesis Rate and Hyaluronan Product Size Are Independent Functions Differentially Affected by Mutations in a Conserved Tandem B-X7-B Motif. Glycobiology 2017, 27, 154−164. (821) Vigetti, D.; Ori, M.; Viola, M.; Genasetti, A.; Karousou, E.; Rizzi, M.; Pallotti, F.; Nardi, I.; Hascall, V. C.; De Luca, G.; et al. Molecular Cloning and Characterization of Udp-Glucose Dehydrogenase from the Amphibian Xenopus Laevis and Its Involvement in Hyaluronan Synthesis. J. Biol. Chem. 2006, 281, 8254−8263. (822) Tian, X.; Azpurua, J.; Hine, C.; Vaidya, A.; MyakishevRempel, M.; Ablaeva, J.; Mao, Z.; Nevo, E.; Gorbunova, V.; Seluanov, A. High-Molecular-Mass Hyaluronan Mediates the Cancer Resistance of the Naked Mole Rat. Nature 2013, 499, 346−349. (823) Csoka, A. B.; Frost, G. I.; Stern, R. The Six HyaluronidaseLike Genes in the Human and Mouse Genomes. Matrix Biol. 2001, 20, 499−508. (824) Stern, R.; Kogan, G.; Jedrzejas, M. J.; Soltés, L. The Many Ways to Cleave Hyaluronan. Biotechnol. Adv. 2007, 25, 537−557. (825) Dumaresq-Doiron, K.; Edjekouane, L.; Orimoto, A. M.; Yoffou, P. H.; Gushulak, L.; Triggs-Raine, B.; Carmona, E. Hyal-1 but Not Hyal-3 Deficiency Has an Impact on Ovarian Folliculogenesis and Female Fertility by Altering the Follistatin/Activin/Smad3 Pathway and the Apoptotic Process. J. Cell. Physiol. 2012, 227, 1911−1922. (826) Cherr, G. N.; Yudin, A. I.; Overstreet, J. W. The Dual Functions of Gpi-Anchored Ph-20: Hyaluronidase and Intracellular Signaling. Matrix Biol. 2001, 20, 515−525. (827) Girish, K. S.; Kemparaju, K. The Magic Glue Hyaluronan and Its Eraser Hyaluronidase: A Biological Overview. Life Sci. 2007, 80, 1921−1943. (828) Bourguignon, L. Y.; Singleton, P. A.; Diedrich, F. HyaluronanCd44 Interaction with Rac1-Dependent Protein Kinase N-Gamma Promotes Phospholipase Cgamma1 Activation, Ca(2+) Signaling, and Cortactin-Cytoskeleton Function Leading to Keratinocyte Adhesion and Differentiation. J. Biol. Chem. 2004, 279, 29654−29669. (829) Erickson, M.; Stern, R. Chain Gangs: New Aspects of Hyaluronan Metabolism. Biochem. Res. Int. 2012, 2012, 893947. (830) Infante, J. R.; Korn, R. L.; Rosen, L. S.; LoRusso, P.; Dychter, S. S.; Zhu, J.; Maneval, D. C.; Jiang, P.; Shepard, H. M.; Frost, G.; et al. Phase 1 Trials of Pegylated Recombinant Human Hyaluronidase Ph20 in Patients with Advanced Solid Tumours. Br. J. Cancer 2018, 118, e3. (831) Yamaguchi, Y.; Yamamoto, H.; Tobisawa, Y.; Irie, F. Tmem2: A Missing Link in Hyaluronan Catabolism Identified? Matrix Biol. 2018, DOI: 10.1016/j.matbio.2018.03.020. (832) Yamamoto, H.; Tobisawa, Y.; Inubushi, T.; Irie, F.; Ohyama, C.; Yamaguchi, Y. A Mammalian Homolog of the Zebrafish Transmembrane Protein 2 (Tmem2) Is the Long-Sought-after CellSurface Hyaluronidase. J. Biol. Chem. 2017, 292, 7304−7313. (833) Nagaoka, A.; Yoshida, H.; Nakamura, S.; Morikawa, T.; Kawabata, K.; Kobayashi, M.; Sakai, S.; Takahashi, Y.; Okada, Y.; Inoue, S. Regulation of Hyaluronan (Ha) Metabolism Mediated by Hybid (Hyaluronan-Binding Protein Involved in Ha Depolymerization, Kiaa1199) and Ha Synthases in Growth Factor-Stimulated Fibroblasts. J. Biol. Chem. 2015, 290, 30910−30923. (834) Zhang, Y.; Jia, S.; Jiang, W. G. Kiaa1199 and Its Biological Role in Human Cancer and Cancer Cells (Review). Oncol. Rep. 2014, 31, 1503−1508. (835) Yoshida, H.; Nagaoka, A.; Komiya, A.; Aoki, M.; Nakamura, S.; Morikawa, T.; Ohtsuki, R.; Sayo, T.; Okada, Y.; Takahashi, Y. Reduction of Hyaluronan and Increased Expression of Hybid BX

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(Kiaa1199) Correlate with Clinical Symptoms in Photoaged Skin. Br. J. Dermatol. 2018, 179,136. (836) McAtee, C. O.; Barycki, J. J.; Simpson, M. A. Emerging Roles for Hyaluronidase in Cancer Metastasis and Therapy. Adv. Cancer Res. 2014, 123, 1−34. (837) Maquart, F. X.; Siméon, A.; Pasco, S.; Monboisse, J. C. Regulation of Cell Activity by the Extracellular Matrix: The Concept of Matrikines. J. Soc. Biol. 1999, 193, 423−428. (838) Vigetti, D.; Deleonibus, S.; Moretto, P.; Bowen, T.; Fischer, J. W.; Grandoch, M.; Oberhuber, A.; Love, D. C.; Hanover, J. A.; Cinquetti, R.; et al. Natural Antisense Transcript for Hyaluronan Synthase 2 (Has2-As1) Induces Transcription of Has2 Via Protein OGlcnacylation. J. Biol. Chem. 2014, 289, 28816−28826. (839) Vigetti, D.; Passi, A. Hyaluronan Synthases Posttranslational Regulation in Cancer. Adv. Cancer Res. 2014, 123, 95−119. (840) Chao, H.; Spicer, A. P. Natural Antisense Mrnas to Hyaluronan Synthase 2 Inhibit Hyaluronan Biosynthesis and Cell Proliferation. J. Biol. Chem. 2005, 280, 27513−27522. (841) Rinn, J. L.; Chang, H. Y. Genome Regulation by Long Noncoding Rnas. Annu. Rev. Biochem. 2012, 81, 145−166. (842) Vigetti, D.; Deleonibus, S.; Viola, M.; Karousou, E.; D’Angelo, M. L.; De Luca, G.; Passi, A. The Long Non-Coding Rna Has2-As1 Enhances the Transcription of Hyaluronan Synthase 2. FASEB J. 2014, 28, 1. DOI: 10.1096/fasebj.28.1_supplement.1005.1 (843) Moretto, P.; Karousou, E.; Viola, M.; Caon, I.; D’Angelo, M. L.; De Luca, G.; Passi, A.; Vigetti, D. Regulation of Hyaluronan Synthesis in Vascular Diseases and Diabetes. J. Diabetes Res. 2015, 2015, 1. (844) Michael, D. R.; Phillips, A. O.; Krupa, A.; Martin, J.; Redman, J. E.; Altaher, A.; Neville, R. D.; Webber, J.; Kim, M. Y.; Bowen, T. The Human Hyaluronan Synthase 2 (Has2) Gene and Its Natural Antisense Rna Exhibit Coordinated Expression in the Renal Proximal Tubular Epithelial Cell. J. Biol. Chem. 2011, 286, 19523−19532. (845) Wang, K. C.; Yang, Y. W.; Liu, B.; Sanyal, A.; CorcesZimmerman, R.; Chen, Y.; Lajoie, B. R.; Protacio, A.; Flynn, R. A.; Gupta, R. A.; et al. A Long Noncoding Rna Maintains Active Chromatin to Coordinate Homeotic Gene Expression. Nature 2011, 472, 120−124. (846) Gupta, R. A.; Shah, N.; Wang, K. C.; Kim, J.; Horlings, H. M.; Wong, D. J.; Tsai, M. C.; Hung, T.; Argani, P.; Rinn, J. L.; et al. Long Non-Coding Rna Hotair Reprograms Chromatin State to Promote Cancer Metastasis. Nature 2010, 464, 1071−1076. (847) Quinodoz, S.; Guttman, M. Long Noncoding Rnas: An Emerging Link between Gene Regulation and Nuclear Organization. Trends Cell Biol. 2014, 24, 651−663. (848) Salmena, L.; Poliseno, L.; Tay, Y.; Kats, L.; Pandolfi, P. P. A Cerna Hypothesis: The Rosetta Stone of a Hidden Rna Language? Cell 2011, 146, 353−358. (849) Lee, D. Y.; Shatseva, T.; Jeyapalan, Z.; Du, W. W.; Deng, Z.; Yang, B. B. A 3′-Untranslated Region (3′utr) Induces Organ Adhesion by Regulating Mir-199a* Functions. PLoS One 2009, 4, e4527. (850) Stern, R.; Asari, A. A.; Sugahara, K. N. Hyaluronan Fragments: An Information-Rich System. Eur. J. Cell Biol. 2006, 85, 699−715. (851) Cyphert, J. M.; Trempus, C. S.; Garantziotis, S. Size Matters: Molecular Weight Specificity of Hyaluronan Effects in Cell Biology. Int. J. Cell Biol. 2015, 2015, 563818. (852) Wu, S. C.; Chen, C. H.; Wang, J. Y.; Lin, Y. S.; Chang, J. K.; Ho, M. L. Hyaluronan Size Alters Chondrogenesis of AdiposeDerived Stem Cells Via the Cd44/Erk/Sox-9 Pathway. Acta Biomater. 2018, 66, 224−237. (853) Camenisch, T. D.; Spicer, A. P.; Brehm-Gibson, T.; Biesterfeldt, J.; Augustine, M. L.; Calabro, A.; Kubalak, S.; Klewer, S. E.; McDonald, J. A. Disruption of Hyaluronan Synthase-2 Abrogates Normal Cardiac Morphogenesis and Hyaluronan-Mediated Transformation of Epithelium to Mesenchyme. J. Clin. Invest. 2000, 106, 349−360. (854) Matsumoto, K.; Li, Y.; Jakuba, C.; Sugiyama, Y.; Sayo, T.; Okuno, M.; Dealy, C. N.; Toole, B. P.; Takeda, J.; Yamaguchi, Y.;

et al. Conditional Inactivation of Has2 Reveals a Crucial Role for Hyaluronan in Skeletal Growth, Patterning, Chondrocyte Maturation and Joint Formation in the Developing Limb. Development 2009, 136, 2825−2835. (855) Wang, Y.; Lauer, M. E.; Anand, S.; Mack, J. A.; Maytin, E. V. Hyaluronan Synthase 2 Protects Skin Fibroblasts against Apoptosis Induced by Environmental Stress. J. Biol. Chem. 2014, 289, 32253− 32265. (856) Goncharova, V.; Serobyan, N.; Iizuka, S.; Schraufstatter, I.; de Ridder, A.; Povaliy, T.; Wacker, V.; Itano, N.; Kimata, K.; Orlovskaja, I. A.; et al. Hyaluronan Expressed by the Hematopoietic Microenvironment Is Required for Bone Marrow Hematopoiesis. J. Biol. Chem. 2012, 287, 25419−25433. (857) Vigetti, D.; Viola, M.; Gornati, R.; Ori, M.; Nardi, I.; Passi, A.; De Luca, G.; Bernardini, G. Molecular Cloning, Genomic Organization and Developmental Expression of the Xenopus Laevis Hyaluronan Synthase 3. Matrix Biol. 2003, 22, 511−517. (858) Bakkers, J.; Kramer, C.; Pothof, J.; Quaedvlieg, N. E.; Spaink, H. P.; Hammerschmidt, M. Has2 Is Required Upstream of Rac1 to Govern Dorsal Migration of Lateral Cells During Zebrafish Gastrulation. Development 2004, 131, 525−537. (859) Chanmee, T.; Ontong, P.; Kimata, K.; Itano, N. Key Roles of Hyaluronan and Its Cd44 Receptor in the Stemness and Survival of Cancer Stem Cells. Front. Oncol. 2015, 5, 180. (860) Ouyang, X.; Panetta, N. J.; Talbott, M. D.; Payumo, A. Y.; Halluin, C.; Longaker, M. T.; Chen, J. K. Hyaluronic Acid Synthesis Is Required for Zebrafish Tail Fin Regeneration. PLoS One 2017, 12, e0171898. (861) Viola, M.; Karousou, E.; D’Angelo, M. L.; Caon, I.; De Luca, G.; Passi, A.; Vigetti, D. Regulated Hyaluronan Synthesis by Vascular Cells. Int. J. Cell Biol. 2015, 2015, 208303. (862) Fischer, J. W. Role of Hyaluronan in Atherosclerosis: Current Knowledge and Open Questions. Matrix Biol. 2018, DOI: 10.1016/ j.matbio.2018.03.003. (863) Nagy, N.; Kuipers, H. F.; Frymoyer, A. R.; Ishak, H. D.; Bollyky, J. B.; Wight, T. N.; Bollyky, P. L. 4-Methylumbelliferone Treatment and Hyaluronan Inhibition as a Therapeutic Strategy in Inflammation, Autoimmunity, and Cancer. Front. Immunol. 2015, 6, 123. (864) Kashima, Y.; Takahashi, M.; Shiba, Y.; Itano, N.; Izawa, A.; Koyama, J.; Nakayama, J.; Taniguchi, S.; Kimata, K.; Ikeda, U. Crucial Role of Hyaluronan in Neointimal Formation after Vascular Injury. PLoS One 2013, 8, e58760. (865) Kurpios, N. A.; Ibañes, M.; Davis, N. M.; Lui, W.; Katz, T.; Martin, J. F.; Izpisúa Belmonte, J. C.; Tabin, C. J. The Direction of Gut Looping Is Established by Changes in the Extracellular Matrix and in Cell:Cell Adhesion. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 8499−8506. (866) de la Motte, C. A. Hyaluronan in Intestinal Homeostasis and Inflammation: Implications for Fibrosis. Am. J. Physiol. Gastrointest. Liver Physiol. 2011, 301, G945−949. (867) Filpa, V.; Bistoletti, M.; Caon, I.; Moro, E.; Grimaldi, A.; Moretto, P.; Baj, A.; Giron, M. C.; Karousou, E.; Viola, M.et al. Changes in Hyaluronan Deposition in the Rat Myenteric Plexus after Experimentally-Induced Colitis. Sci. Rep. 2017, 7, DOI: 10.1038/ s41598-017-18020-7. (868) Neuman, M. G.; Cohen, L. B.; Nanau, R. M. Hyaluronic Acid as a Non-Invasive Biomarker of Liver Fibrosis. Clin. Biochem. 2016, 49, 302−315. (869) Perkins, K. L.; Arranz, A. M.; Yamaguchi, Y.; Hrabetova, S. Brain Extracellular Space, Hyaluronan, and the Prevention of Epileptic Seizures. Rev. Neurosci. 2017, 28, 869−892. (870) Vigetti, D.; Andrini, O.; Clerici, M.; Negrini, D.; Passi, A.; Moriondo, A. Chondroitin Sulfates Act as Extracellular Gating Modifiers on Voltage-Dependent Ion Channels. Cell. Physiol. Biochem. 2008, 22, 137−146. (871) Arranz, A. M.; Perkins, K. L.; Irie, F.; Lewis, D. P.; Hrabe, J.; Xiao, F.; Itano, N.; Kimata, K.; Hrabetova, S.; Yamaguchi, Y. Hyaluronan Deficiency Due to Has3 Knock-out Causes Altered BY

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

Neuronal Activity and Seizures Via Reduction in Brain Extracellular Space. J. Neurosci. 2014, 34, 6164−6176. (872) Olsson, M.; Meadows, J. R.; Truvé, K.; Rosengren Pielberg, G.; Puppo, F.; Mauceli, E.; Quilez, J.; Tonomura, N.; Zanna, G.; Docampo, M. J.; et al. A Novel Unstable Duplication Upstream of Has2 Predisposes to a Breed-Defining Skin Phenotype and a Periodic Fever Syndrome in Chinese Shar-Pei Dogs. PLoS Genet. 2011, 7, e1001332. (873) Aya, K. L.; Stern, R. Hyaluronan in Wound Healing: Rediscovering a Major Player. Wound Repair Regen. 2014, 22, 579− 593. (874) Plaas, A.; Li, J.; Riesco, J.; Das, R.; Sandy, J. D.; Harrison, A. Intraarticular Injection of Hyaluronan Prevents Cartilage Erosion, Periarticular Fibrosis and Mechanical Allodynia and Normalizes Stance Time in Murine Knee Osteoarthritis. Arthritis Res. Ther. 2011, 13, R46. (875) Burdick, J. A.; Prestwich, G. D. Hyaluronic Acid Hydrogels for Biomedical Applications. Adv. Mater. 2011, 23, H41−56. (876) Lauer, M. E.; Dweik, R. A.; Garantziotis, S.; Aronica, M. A. The Rise and Fall of Hyaluronan in Respiratory Diseases. Int. J. Cell Biol. 2015, 2015, 712507. (877) Tighe, R. M.; Garantziotis, S. Hyaluronan Interactions with Innate Immunity in Lung Biology. Matrix Biol. 2018, DOI: 10.1016/ j.matbio.2018.01.027. (878) Chanmee, T.; Ontong, P.; Itano, N. Hyaluronan: A Modulator of the Tumor Microenvironment. Cancer Lett. 2016, 375, 20−30. (879) Udabage, L.; Brownlee, G. R.; Nilsson, S. K.; Brown, T. J. The over-Expression of Has2, Hyal-2 and Cd44 Is Implicated in the Invasiveness of Breast Cancer. Exp. Cell Res. 2005, 310, 205−217. (880) Edelman, R.; Assaraf, Y. G.; Levitzky, I.; Shahar, T.; Livney, Y. D. Hyaluronic Acid-Serum Albumin Conjugate-Based Nanoparticles for Targeted Cancer Therapy. Oncotarget 2017, 8, 24337−24353. (881) Oikari, S.; Kettunen, T.; Tiainen, S.; Häyrinen, J.; Masarwah, A.; Sudah, M.; Sutela, A.; Vanninen, R.; Tammi, M.; Auvinen, P. UdpSugar Accumulation Drives Hyaluronan Synthesis in Breast Cancer. Matrix Biol. 2018, 67, 63−74. (882) Viola, M.; Vigetti, D.; Karousou, E.; D’Angelo, M. L.; Caon, I.; Moretto, P.; De Luca, G.; Passi, A. Biology and Biotechnology of Hyaluronan. Glycoconjugate J. 2015, 32, 93−103. (883) Piperigkou, Z.; Karamanou, K.; Engin, A. B.; Gialeli, C.; Docea, A. O.; Vynios, D. H.; Pavao, M. S.; Golokhvast, K. S.; Shtilman, M. I.; Argiris, A.; et al. Emerging Aspects of Nanotoxicology in Health and Disease: From Agriculture and Food Sector to Cancer Therapeutics. Food Chem. Toxicol. 2016, 91, 42−57. (884) Tsai, S. W.; Fang, J. F.; Yang, C. L.; Chen, J. H.; Su, L. T.; Jan, S. H. Preparation and Evaluation of a Hyaluronate-Collagen Film for Preventing Post-Surgical Adhesion. J. Int. Med. Res. 2005, 33, 68−76. (885) Tamer, T. M. Hyaluronan and Synovial Joint: Function, Distribution and Healing. Interdiscip. Toxicol. 2013, 6, 111−125. (886) Morris, E. R.; Rees, D. A.; Welsh, E. J. Conformation and Dynamic Interactions in Hyaluronate Solutions. J. Mol. Biol. 1980, 138, 383−400. (887) Cowman, M. K.; Matsuoka, S. Experimental Approaches to Hyaluronan Structure. Carbohydr. Res. 2005, 340, 791−809. (888) Soltés, L.; Mendichi, R.; Kogan, G.; Schiller, J.; Stankovska, M.; Arnhold, J. Degradative Action of Reactive Oxygen Species on Hyaluronan. Biomacromolecules 2006, 7, 659−668. (889) Simkovic, I. Unexplored Possibilities of All-Polysaccharide Composites. Carbohydr. Polym. 2013, 95, 697−715. (890) Itano, N. Simple Primary Structure, Complex Turnover Regulation and Multiple Roles of Hyaluronan. J. Biochem. 2008, 144, 131−137. (891) Voigt, J.; Driver, V. R. Hyaluronic Acid Derivatives and Their Healing Effect on Burns, Epithelial Surgical Wounds, and Chronic Wounds: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. Wound Repair Regen. 2012, 20, 317−331. (892) Tamer, T. M.; Collins, M. N.; Valachová, K.; Hassan, M. A.; Omer, A. M.; Mohy-Eldin, M. S.; Š vík, K.; Jurčík, R.; Ondruška, L.;

Biró, C.et al. Mitoq Loaded Chitosan-Hyaluronan Composite Membranes for Wound Healing. Materials 2018, 11,569. (893) Damodarasamy, M.; Johnson, R. S.; Bentov, I.; MacCoss, M. J.; Vernon, R. B.; Reed, M. J. Hyaluronan Enhances Wound Repair and Increases Collagen Iii in Aged Dermal Wounds. Wound Repair Regen. 2014, 22, 521−526. (894) Jiang, Z.; Dong, X.; Yan, X.; Liu, Y.; Zhang, L.; Sun, Y. Nanogels of Dual Inhibitor-Modified Hyaluronic Acid Function as a Potent Inhibitor of Amyloid B-Protein Aggregation and Cytotoxicity. Sci. Rep. 2018, 8, 3505. (895) Hsu, M. F.; Tyan, Y. S.; Chien, Y. C.; Lee, M. W. Hyaluronic Acid-Based Nano-Sized Drug Carrier-Containing Gellan Gum Microspheres as Potential Multifunctional Embolic Agent. Sci. Rep. 2018, 8, 731. (896) Darr, A.; Calabro, A. Synthesis and Characterization of Tyramine-Based Hyaluronan Hydrogels. J. Mater. Sci.: Mater. Med. 2009, 20, 33−44. (897) Gariboldi, S.; Palazzo, M.; Zanobbio, L.; Selleri, S.; Sommariva, M.; Sfondrini, L.; Cavicchini, S.; Balsari, A.; Rumio, C. Low Molecular Weight Hyaluronic Acid Increases the Self-Defense of Skin Epithelium by Induction of Beta-Defensin 2 Via Tlr2 and Tlr4. J. Immunol. 2008, 181, 2103−2110. (898) Manuskiatti, W.; Maibach, H. I. Hyaluronic Acid and Skin: Wound Healing and Aging. Int. J. Dermatol. 1996, 35, 539−544. (899) Maytin, E. V. Hyaluronan: More Than Just a Wrinkle Filler. Glycobiology 2016, 26, 553−559. (900) Fallacara, A.; Manfredini, S.; Durini, E.; Vertuani, S. Hyaluronic Acid Fillers in Soft Tissue Regeneration. Facial Plast. Surg. 2017, 33, 87−96. (901) Duranti, F.; Salti, G.; Bovani, B.; Calandra, M.; Rosati, M. L. Injectable Hyaluronic Acid Gel for Soft Tissue Augmentation. A Clinical and Histological Study. Dermatol. Surg. 1998, 24, 1317−1325. (902) Narins, R. S.; Brandt, F.; Leyden, J.; Lorenc, Z. P.; Rubin, M.; Smith, S. A Randomized, Double-Blind, Multicenter Comparison of the Efficacy and Tolerability of Restylane Versus Zyplast for the Correction of Nasolabial Folds. Dermatol. Surg. 2003, 29, 588−595. (903) Varki, A.; Cummings, R. D.; Aebi, M.; Packer, N. H.; Seeberger, P. H.; Esko, J. D.; Stanley, P.; Hart, G.; Darvill, A.; Kinoshita, T.; et al. Symbol Nomenclature for Graphical Representations of Glycans. Glycobiology 2015, 25, 1323−1324. (904) Billings, P. C.; Pacifici, M. Interactions of Signaling Proteins, Growth Factors and Other Proteins with Heparan Sulfate: Mechanisms and Mysteries. Connect. Tissue Res. 2015, 56, 272−280. (905) Vorup-Jensen, T.; Chi, L.; Gjelstrup, L. C.; Jensen, U. B.; Jewett, C. A.; Xie, C.; Shimaoka, M.; Linhardt, R. J.; Springer, T. A. Binding between the Integrin Alphaxbeta2 (Cd11c/Cd18) and Heparin. J. Biol. Chem. 2007, 282, 30869−30877. (906) Faye, C.; Chautard, E.; Olsen, B. R.; Ricard-Blum, S. The First Draft of the Endostatin Interaction Network. J. Biol. Chem. 2009, 284, 22041−22047. (907) Faye, C.; Moreau, C.; Chautard, E.; Jetne, R.; Fukai, N.; Ruggiero, F.; Humphries, M. J.; Olsen, B. R.; Ricard-Blum, S. Molecular Interplay between Endostatin, Integrins, and Heparan Sulfate. J. Biol. Chem. 2009, 284, 22029−22040. (908) Rider, C. C.; Mulloy, B. Heparin, Heparan Sulphate and the Tgf-Beta Cytokine Superfamily. Molecules 2017, 22 (5), 73. (909) Bartlett, A. H.; Park, P. W. Proteoglycans in Host-Pathogen Interactions: Molecular Mechanisms and Therapeutic Implications. Expert Rev. Mol. Med. 2010, 12, e5. (910) Peysselon, F.; Ricard-Blum, S. Heparin-Protein Interactions: From Affinity and Kinetics to Biological Roles. Application to an Interaction Network Regulating Angiogenesis. Matrix Biol. 2014, 35, 73−81. (911) Kamhi, E.; Joo, E. J.; Dordick, J. S.; Linhardt, R. J. Glycosaminoglycans in Infectious Disease. Biol. Rev. Camb. Philos. Soc. 2013, 88, 928−943. (912) Aquino, R. S.; Park, P. W. Glycosaminoglycans and Infection. Front. Biosci., Landmark Ed. 2016, 21, 1260−1277. BZ

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(913) Kim, S. Y.; Zhao, J.; Liu, X.; Fraser, K.; Lin, L.; Zhang, X.; Zhang, F.; Dordick, J. S.; Linhardt, R. J. Interaction of Zika Virus Envelope Protein with Glycosaminoglycans. Biochemistry 2017, 56, 1151−1162. (914) Launay, G.; Salza, R.; Multedo, D.; Thierry-Mieg, N.; RicardBlum, S. Matrixdb, the Extracellular Matrix Interaction Database: Updated Content, a New Navigator and Expanded Functionalities. Nucleic Acids Res. 2015, 43, D321−327. (915) Apweiler, R.; Bairoch, A.; Wu, C. H.; Barker, W. C.; Boeckmann, B.; Ferro, S.; Gasteiger, E.; Huang, H.; Lopez, R.; Magrane, M.; et al. Uniprot: The Universal Protein Knowledgebase. Nucleic Acids Res. 2004, 32, D115−D119. (916) Meldal, B. H.; Forner-Martinez, O.; Costanzo, M. C.; Dana, J.; Demeter, J.; Dumousseau, M.; Dwight, S. S.; Gaulton, A.; Licata, L.; Melidoni, A. N.; et al. The Complex Portal–an Encyclopaedia of Macromolecular Complexes. Nucleic Acids Res. 2015, 43, D479−484. (917) Shannon, P.; Markiel, A.; Ozier, O.; Baliga, N. S.; Wang, J. T.; Ramage, D.; Amin, N.; Schwikowski, B.; Ideker, T. Cytoscape: A Software Environment for Integrated Models of Biomolecular Interaction Networks. Genome Res. 2003, 13, 2498−2504. (918) Meneghetti, M. C.; Hughes, A. J.; Rudd, T. R.; Nader, H. B.; Powell, A. K.; Yates, E. A.; Lima, M. A. Heparan Sulfate and Heparin Interactions with Proteins. J. R. Soc., Interface 2015, 12, 0589. (919) Ori, A.; Wilkinson, M. C.; Fernig, D. G. A Systems Biology Approach for the Investigation of the Heparin/Heparan Sulfate Interactome. J. Biol. Chem. 2011, 286, 19892−19904. (920) Nikitovic, D.; Kouvidi, K.; Kavasi, R. M.; Berdiaki, A.; Tzanakakis, G. N. Hyaluronan/Hyaladherins - a Promising Axis for Targeted Drug Delivery in Cancer. Curr. Drug Delivery 2016, 13, 500−511. (921) Conrad, A. H.; Zhang, Y.; Tasheva, E. S.; Conrad, G. W. Proteomic Analysis of Potential Keratan Sulfate, Chondroitin Sulfate a, and Hyaluronic Acid Molecular Interactions. Invest. Ophthalmol. Visual Sci. 2010, 51, 4500−4515. (922) Evanko, S. P.; Parks, W. T.; Wight, T. N. Intracellular Hyaluronan in Arterial Smooth Muscle Cells: Association with Microtubules, Rhamm, and the Mitotic Spindle. J. Histochem. Cytochem. 2004, 52, 1525−1535. (923) Stewart, M. D.; Sanderson, R. D. Heparan Sulfate in the Nucleus and Its Control of Cellular Functions. Matrix Biol. 2014, 35, 56−59. (924) Stein, G. S.; Roberts, R. M.; Davis, J. L.; Head, W. J.; Stein, J. L.; Thrall, C. L.; Van Veen, J.; Welch, D. W. Are Glycoproteins and Glycosaminoglycans Components of the Eukaryotic Genome? Nature 1975, 258, 639−641. (925) Hiscock, D. R.; Yanagishita, M.; Hascall, V. C. Nuclear Localization of Glycosaminoglycans in Rat Ovarian Granulosa Cells. J. Biol. Chem. 1994, 269, 4539−4546. (926) Sun, C.; Marcello, M.; Li, Y.; Mason, D.; Levy, R.; Fernig, D. G. Selectivity in Glycosaminoglycan Binding Dictates the Distribution and Diffusion of Fibroblast Growth Factors in the Pericellular Matrix. Open Biol. 2016, 6, 150277. (927) Saesen, E.; Sarrazin, S.; Laguri, C.; Sadir, R.; Maurin, D.; Thomas, A.; Imberty, A.; Lortat-Jacob, H. Insights into the Mechanism by Which Interferon-Gamma Basic Amino Acid Clusters Mediate Protein Binding to Heparan Sulfate. J. Am. Chem. Soc. 2013, 135, 9384−9390. (928) Sage, J.; Mallevre, F.; Barbarin-Costes, F.; Samsonov, S. A.; Gehrcke, J. P.; Pisabarro, M. T.; Perrier, E.; Schnebert, S.; Roget, A.; Livache, T.; et al. Binding of Chondroitin 4-Sulfate to Cathepsin S Regulates Its Enzymatic Activity. Biochemistry 2013, 52, 6487−6498. (929) Dyer, D. P.; Salanga, C. L.; Volkman, B. F.; Kawamura, T.; Handel, T. M. The Dependence of Chemokine-Glycosaminoglycan Interactions on Chemokine Oligomerization. Glycobiology 2015, 26, 312−326. (930) Vallet, S. D.; Deddens, L.; Vonarburg, A.; Salza, R.; Faye, C.; Aranyos, A.; Thierry-Mieg, N.; Ricard-Blum, S. In Handbook of Surface Plasmon Resonance (2); The Royal Society of Chemistry, 2017, DOI: 10.1039/9781788010283-00398.

(931) Beurdeley, M.; Spatazza, J.; Lee, H. H.; Sugiyama, S.; Bernard, C.; Di Nardo, A. A.; Hensch, T. K.; Prochiantz, A. Otx2 Binding to Perineuronal Nets Persistently Regulates Plasticity in the Mature Visual Cortex. J. Neurosci. 2012, 32, 9429−9437. (932) Corredor, M.; Bonet, R.; Moure, A.; Domingo, C.; Bujons, J.; Alfonso, I.; Perez, Y.; Messeguer, A. Cationic Peptides and Peptidomimetics Bind Glycosaminoglycans as Potential Sema3a Pathway Inhibitors. Biophys. J. 2016, 110, 1291−1303. (933) Dutta, A. K.; Rosgen, J.; Rajarathnam, K. Using Isothermal Titration Calorimetry to Determine Thermodynamic Parameters of Protein-Glycosaminoglycan Interactions. Methods Mol. Biol. 2015, 1229, 315−324. (934) Ulbricht, D.; Oertwig, K.; Arnsburg, K.; Saalbach, A.; Pippel, J.; Strater, N.; Heiker, J. T. Basic Residues of Beta-Sheet a Contribute to Heparin Binding and Activation of Vaspin (Serpin A12). J. Biol. Chem. 2017, 292, 994−1004. (935) Minsky, B. B.; Antoni, C. H.; Boehm, H. Controlled Immobilization Strategies to Probe Short Hyaluronan-Protein Interactions. Sci. Rep. 2016, 6, 21608. (936) Salanga, C. L.; Handel, T. M. Chemokine Oligomerization and Interactions with Receptors and Glycosaminoglycans: The Role of Structural Dynamics in Function. Exp. Cell Res. 2011, 317, 590− 601. (937) Imberty, A.; Lortat-Jacob, H.; Perez, S. Structural View of Glycosaminoglycan-Protein Interactions. Carbohydr. Res. 2007, 342, 430−439. (938) Pomin, V. H.; Mulloy, B. Glycosaminoglycans and Proteoglycans. Pharmaceuticals 2018, 11, 27. (939) Vives, R. R.; Crublet, E.; Andrieu, J. P.; Gagnon, J.; Rousselle, P.; Lortat-Jacob, H. A Novel Strategy for Defining Critical Amino Acid Residues Involved in Protein/Glycosaminoglycan Interactions. J. Biol. Chem. 2004, 279, 54327−54333. (940) Ori, A.; Free, P.; Courty, J.; Wilkinson, M. C.; Fernig, D. G. Identification of Heparin-Binding Sites in Proteins by Selective Labeling. Mol. Cell. Proteomics 2009, 8, 2256−2265. (941) Mobius, K.; Nordsieck, K.; Pichert, A.; Samsonov, S. A.; Thomas, L.; Schiller, J.; Kalkhof, S.; Teresa Pisabarro, M.; BeckSickinger, A. G.; Huster, D. Investigation of Lysine Side Chain Interactions of Interleukin-8 with Heparin and Other Glycosaminoglycans Studied by a Methylation-Nmr Approach. Glycobiology 2013, 23, 1260−1269. (942) Seffouh, A.; Milz, F.; Przybylski, C.; Laguri, C.; Oosterhof, A.; Bourcier, S.; Sadir, R.; Dutkowski, E.; Daniel, R.; van Kuppevelt, T. H.; et al. Hsulf Sulfatases Catalyze Processive and Oriented 6-ODesulfation of Heparan Sulfate That Differentially Regulates Fibroblast Growth Factor Activity. FASEB J. 2013, 27, 2431−2439. (943) Pichert, A.; Samsonov, S. A.; Theisgen, S.; Thomas, L.; Baumann, L.; Schiller, J.; Beck-Sickinger, A. G.; Huster, D.; Pisabarro, M. T. Characterization of the Interaction of Interleukin-8 with Hyaluronan, Chondroitin Sulfate, Dermatan Sulfate and Their Sulfated Derivatives by Spectroscopy and Molecular Modeling. Glycobiology 2012, 22, 134−145. (944) Laguri, C.; Sapay, N.; Simorre, J. P.; Brutscher, B.; Imberty, A.; Gans, P.; Lortat-Jacob, H. 13c-Labeled Heparan Sulfate Analogue as a Tool to Study Protein/Heparan Sulfate Interactions by Nmr Spectroscopy: Application to the Cxcl12alpha Chemokine. J. Am. Chem. Soc. 2011, 133, 9642−9645. (945) Hofmann, T.; Samsonov, S. A.; Pichert, A.; Lemmnitzer, K.; Schiller, J.; Huster, D.; Pisabarro, M. T.; von Bergen, M.; Kalkhof, S. Structural Analysis of the Interleukin-8/Glycosaminoglycan Interactions by Amide Hydrogen/Deuterium Exchange Mass Spectrometry. Methods 2015, 89, 45−53. (946) Capila, I.; Linhardt, R. J. Heparin-Protein Interactions. Angew. Chem., Int. Ed. 2002, 41, 390−412. (947) Sarkar, A.; Desai, U. R. A Simple Method for Discovering Druggable, Specific Glycosaminoglycan-Protein Systems. Elucidation of Key Principles from Heparin/Heparan Sulfate-Binding Proteins. PLoS One 2015, 10, e0141127. CA

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(948) Luo, Y.; Lu, W.; Mohamedali, K. A.; Jang, J. H.; Jones, R. B.; Gabriel, J. L.; Kan, M.; McKeehan, W. L. The Glycine Box: A Determinant of Specificity for Fibroblast Growth Factor. Biochemistry 1998, 37, 16506−16515. (949) Torrent, M.; Nogues, M. V.; Andreu, D.; Boix, E. The ″Cpc Clip Motif″: A Conserved Structural Signature for Heparin-Binding Proteins. PLoS One 2012, 7, e42692. (950) Dicker, K. T.; Gurski, L. A.; Pradhan-Bhatt, S.; Witt, R. L.; Farach-Carson, M. C.; Jia, X. Hyaluronan: A Simple Polysaccharide with Diverse Biological Functions. Acta Biomater. 2014, 10, 1558− 1570. (951) Amemiya, K.; Nakatani, T.; Saito, A.; Suzuki, A.; Munakata, H. Hyaluronan-Binding Motif Identified by Panning a Random Peptide Display Library. Biochim. Biophys. Acta, Gen. Subj. 2005, 1724, 94−99. (952) Yang, B.; Yang, B. L.; Savani, R. C.; Turley, E. A. Identification of a Common Hyaluronan Binding Motif in the Hyaluronan Binding Proteins Rhamm, Cd44 and Link Protein. EMBO J. 1994, 13, 286− 296. (953) 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. (954) Sankaranarayanan, N. V.; Strebel, T. R.; Boothello, R. S.; Sheerin, K.; Raghuraman, A.; Sallas, F.; Mosier, P. D.; Watermeyer, N. D.; Oscarson, S.; Desai, U. R. A Hexasaccharide Containing Rare 2-OSulfate-Glucuronic Acid Residues Selectively Activates Heparin Cofactor Ii. Angew. Chem., Int. Ed. 2017, 56, 2312−2317. (955) El Masri, R.; Seffouh, A.; Lortat-Jacob, H.; Vives, R. R. The ″in and out″ of Glucosamine 6-O-Sulfation: The 6th Sense of Heparan Sulfate. Glycoconjugate J. 2017, 34, 285−298. (956) Vivès, R. R.; Seffouh, A.; Lortat-Jacob, H. Post-Synthetic Regulation of Hs Structure: The Yin and Yang of the Sulfs in Cancer. Front. Oncol. 2014, 3, 331. (957) Ricard-Blum, S.; Feraud, O.; Lortat-Jacob, H.; Rencurosi, A.; Fukai, N.; Dkhissi, F.; Vittet, D.; Imberty, A.; Olsen, B. R.; van der Rest, M. Characterization of Endostatin Binding to Heparin and Heparan Sulfate by Surface Plasmon Resonance and Molecular Modeling: Role of Divalent Cations. J. Biol. Chem. 2004, 279, 2927− 2936. (958) Samsonov, S. A.; Pisabarro, M. T. Computational Analysis of Interactions in Structurally Available Protein-Glycosaminoglycan Complexes. Glycobiology 2016, 26, 850−861. (959) Babik, S.; Samsonov, S. A.; Pisabarro, M. T. Computational Drill Down on Fgf1-Heparin Interactions through Methodological Evaluation. Glycoconjugate J. 2017, 34, 427−440. (960) Takaoka, T.; Mori, K.; Okimoto, N.; Neya, S.; Hoshino, T. Prediction of the Structure of Complexes Comprised of Proteins and Glycosaminoglycans Using Docking Simulation and Cluster Analysis. J. Chem. Theory Comput. 2007, 3, 2347−2356. (961) Ballut, L.; Sapay, N.; Chautard, E.; Imberty, A.; Ricard-Blum, S. Mapping of Heparin/Heparan Sulfate Binding Sites on Alphavbeta3 Integrin by Molecular Docking. J. Mol. Recognit. 2013, 26, 76−85. (962) Samsonov, S. A.; Teyra, J.; Pisabarro, M. T. Docking Glycosaminoglycans to Proteins: Analysis of Solvent Inclusion. J. Comput.-Aided Mol. Des. 2011, 25, 477−489. (963) Sapay, N.; Cabannes, E.; Petitou, M.; Imberty, A. Molecular Modeling of the Interaction between Heparan Sulfate and Cellular Growth Factors: Bringing Pieces Together. Glycobiology 2011, 21, 1181−1193. (964) Samsonov, S. A.; Pisabarro, M. T. Importance of Idoa and Idoa(2s) Ring Conformations in Computational Studies of Glycosaminoglycan-Protein Interactions. Carbohydr. Res. 2013, 381, 133− 137. (965) Samsonov, S. A.; Gehrcke, J. P.; Pisabarro, M. T. Flexibility and Explicit Solvent in Molecular-Dynamics-Based Docking of Protein-Glycosaminoglycan Systems. J. Chem. Inf. Model. 2014, 54, 582−592.

(966) Griffith, A. R.; Rogers, C. J.; Miller, G. M.; Abrol, R.; HsiehWilson, L. C.; Goddard, W. A., 3rd. Predicting Glycosaminoglycan Surface Protein Interactions and Implications for Studying Axonal Growth. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 13697−13702. (967) Sankaranarayanan, N. V.; Sarkar, A.; Desai, U. R.; Mosier, P. D. Designing ″High-Affinity, High-Specificity″ Glycosaminoglycan Sequences through Computerized Modeling. Methods Mol. Biol. 2015, 1229, 289−314. (968) Bathe, M.; Rutledge, G. C.; Grodzinsky, A. J.; Tidor, B. A Coarse-Grained Molecular Model for Glycosaminoglycans: Application to Chondroitin, Chondroitin Sulfate, and Hyaluronic Acid. Biophys. J. 2005, 88, 3870−3887. (969) Samsonov, S. A.; Bichmann, L.; Pisabarro, M. T. CoarseGrained Model of Glycosaminoglycans. J. Chem. Inf. Model. 2015, 55, 114−124. (970) Sattelle, B. M.; Shakeri, J.; Cliff, M. J.; Almond, A. Proteoglycans and Their Heterogeneous Glycosaminoglycans at the Atomic Scale. Biomacromolecules 2015, 16, 951−961. (971) Turnbull, J. E. Heparan Sulfate Glycomics: Towards Systems Biology Strategies. Biochem. Soc. Trans. 2010, 38, 1356−1360. (972) Puvirajesinghe, T. M.; Turnbull, J. E. Glycoarray Technologies: Deciphering Interactions from Proteins to Live Cell Responses. Microarrays 2016, 5, 3. (973) Mercey, E.; Sadir, R.; Maillart, E.; Roget, A.; Baleux, F.; Lortat-Jacob, H.; Livache, T. Polypyrrole Oligosaccharide Array and Surface Plasmon Resonance Imaging for the Measurement of Glycosaminoglycan Binding Interactions. Anal. Chem. 2008, 80, 3476−3482. (974) Rogers, C. J.; Clark, P. M.; Tully, S. E.; Abrol, R.; Garcia, K. C.; Goddard, W. A., 3rd; Hsieh-Wilson, L. C. Elucidating Glycosaminoglycan-Protein-Protein Interactions Using Carbohydrate Microarray and Computational Approaches. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 9747−9752. (975) Dulaney, S. B.; Xu, Y.; Wang, P.; Tiruchinapally, G.; Wang, Z.; Kathawa, J.; El-Dakdouki, M. H.; Yang, B.; Liu, J.; Huang, X. Divergent Synthesis of Heparan Sulfate Oligosaccharides. J. Org. Chem. 2015, 80, 12265−12279. (976) Puvirajesinghe, T. M.; Ahmed, Y. A.; Powell, A. K.; Fernig, D. G.; Guimond, S. E.; Turnbull, J. E. Array-Based Functional Screening of Heparin Glycans. Chem. Biol. 2012, 19, 553−558. (977) Fatoux-Ardore, M.; Peysselon, F.; Weiss, A.; Bastien, P.; Pratlong, F.; Ricard-Blum, S. Large-Scale Investigation of Leishmania Interaction Networks with Host Extracellular Matrix by Surface Plasmon Resonance Imaging. Infect. Immun. 2014, 82, 594−606. (978) Otto, D. M.; Campanero-Rhodes, M. A.; Karamanska, R.; Powell, A. K.; Bovin, N.; Turnbull, J. E.; Field, R. A.; Blackburn, J.; Feizi, T.; Crocker, P. R. An Expression System for Screening of Proteins for Glycan and Protein Interactions. Anal. Biochem. 2011, 411, 261−270. (979) Yang, J.; Hsieh, P. H.; Liu, X.; Zhou, W.; Zhang, X.; Zhao, J.; Xu, Y.; Zhang, F.; Linhardt, R. J.; Liu, J. Construction and Characterisation of a Heparan Sulphate Heptasaccharide Microarray. Chem. Commun. (Cambridge, U. K.) 2017, 53, 1743−1746. (980) Salza, R.; Lethias, C.; Ricard-Blum, S. The Multimerization State of the Amyloid-Beta42 Amyloid Peptide Governs Its Interaction Network with the Extracellular Matrix. J. Alzheimer's Dis. 2017, 56, 991−1005. (981) Hahm, H. S.; Schlegel, M. K.; Hurevich, M.; Eller, S.; Schuhmacher, F.; Hofmann, J.; Pagel, K.; Seeberger, P. H. Automated Glycan Assembly Using the Glyconeer 2.1 Synthesizer. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, E3385−E3389. (982) Powell, A. K.; Zhi, Z. L.; Turnbull, J. E. Saccharide Microarrays for High-Throughput Interrogation of Glycan-Protein Binding Interactions. Methods Mol. Biol. 2009, 534, 313−329. (983) Liu, Y.; Childs, R. A.; Palma, A. S.; Campanero-Rhodes, M. A.; Stoll, M. S.; Chai, W.; Feizi, T. Neoglycolipid-Based Oligosaccharide Microarray System: Preparation of Ngls and Their Noncovalent Immobilization on Nitrocellulose-Coated Glass Slides for Microarray Analyses. Methods Mol. Biol. 2012, 808, 117−136. CB

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX

Chemical Reviews

Review

(984) Fukui, S.; Feizi, T.; Galustian, C.; Lawson, A. M.; Chai, W. Oligosaccharide Microarrays for High-Throughput Detection and Specificity Assignments of Carbohydrate-Protein Interactions. Nat. Biotechnol. 2002, 20, 1011−1017. (985) Hsiao, F. S.; Sutandy, F. R.; Syu, G. D.; Chen, Y. W.; Lin, J. M.; Chen, C. S. Systematic Protein Interactome Analysis of Glycosaminoglycans Revealed Ycbs as a Novel Bacterial Virulence Factor. Sci. Rep. 2016, 6, 28425. (986) Ricard-Blum, S.; Lisacek, F. Glycosaminoglycanomics: Where We Are. Glycoconjugate J. 2017, 34, 339−349. (987) Ricard-Blum, S. Glycosaminoglycans: Major Biological Players. Glycoconjugate J. 2017, 34, 275−276. (988) Orchard, S.; Kerrien, S.; Abbani, S.; Aranda, B.; Bhate, J.; Bidwell, S.; Bridge, A.; Briganti, L.; Brinkman, F. S.; Cesareni, G.; et al. Protein Interaction Data Curation: The International Molecular Exchange (Imex) Consortium. Nat. Methods 2012, 9, 345−350. (989) Nunes, Q. M.; Mournetas, V.; Lane, B.; Sutton, R.; Fernig, D. G.; Vasieva, O. The Heparin-Binding Protein Interactome in Pancreatic Diseases. Pancreatology 2013, 13, 598−604. (990) Lackey, B.; Nunes, Q. M.; Higham, S. M.; Fernig, D. G.; Valappil, S. P. In Silico Analyses of Heparin Binding Proteins Expression in Human Periodontal Tissues. BMC Res. Notes 2016, 9, 53. (991) Pathan, M.; Keerthikumar, S.; Ang, C. S.; Gangoda, L.; Quek, C. Y.; Williamson, N. A.; Mouradov, D.; Sieber, O. M.; Simpson, R. J.; Salim, A.; et al. Funrich: An Open Access Standalone Functional Enrichment and Interaction Network Analysis Tool. Proteomics 2015, 15, 2597−2601. (992) Huang, d. W.; Sherman, B. T.; Lempicki, R. A. Systematic and Integrative Analysis of Large Gene Lists Using David Bioinformatics Resources. Nat. Protoc. 2009, 4, 44−57. (993) Roper, J. A.; Williamson, R. C.; Bass, M. D. Syndecan and Integrin Interactomes: Large Complexes in Small Spaces. Curr. Opin. Struct. Biol. 2012, 22, 583−590.

CC

DOI: 10.1021/acs.chemrev.8b00354 Chem. Rev. XXXX, XXX, XXX−XXX