Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Jun 17.
Published in final edited form as: Exp Neurol. 2013 Nov 4;251:39–46. doi: 10.1016/j.expneurol.2013.10.017

Inhibitors of myelination: ECM changes, CSPGs and PTPs

Danielle E Harlow a,b, Wendy B Macklin a,b,*
PMCID: PMC4060786  NIHMSID: NIHMS589857  PMID: 24200549

Abstract

After inflammation-induced demyelination, such as in the disease multiple sclerosis, endogenous remyelination often fails. However, in animal models of demyelination induced with toxins, remyelination can be quite robust. A significant difference between inflammation-induced and toxin-induced demyelination is the response of local cells within the lesion, including astrocytes, oligodendrocytes, microglia/macrophages, and NG2+ cells, which respond to inflammatory stimuli with increased extracellular matrix (ECM) protein and chondroitin sulfate proteoglycan (CSPG) production and deposition. Here, we summarize current knowledge of ECM changes in demyelinating lesions, as well as oligodendrocyte responses to aberrant ECM proteins and CSPGs after various types of demyelinating insults. The discovery that CSPGs act through the receptor protein tyrosine phosphatase sigma (PTPσ) and the Rho-ROCK pathway to inhibit oligodendrocyte process extension and myelination, but not oligodendrocyte differentiation (Pendleton et al., Experimental Neurology (2013) vol. 247, pp. 113-121), highlights the need to better understand the ECM changes that accompany demyelination and their influence on oligodendrocytes and effective remyelination.

Introduction

A hallmark of central nervous system (CNS) injury is the activation and proliferation of local glial cells, including microglia, astrocytes, and oligodendrocytes. Reactive glial cells, in particular astrocytes and microglia, contribute to formation of the so-called ‘glial scar’, by depositing extracellular matrix proteins and upregulating molecules that are often inhibitory to regeneration (Fitch and Silver, 2008; Galtrey et al., 2008; Morgenstern et al., 2002; Rhodes and Fawcett, 2004; Sherman and Back, 2008). High levels of chondroitin sulfate proteoglycans (CSPGs) are present in the scar after many types of CNS insults including spinal cord injury (SCI; Jones et al., 2003; Lemons et al., 1999; McTigue et al., 2001; Tang et al., 2003), epilepsy (Kurazono et al., 2001; Okamoto et al., 2003), Alzheimer's disease (DeWitt et al., 1993; Snow et al., 1988; 1990), Parkinson's disease (DeWitt et al., 1994), stroke (Carmichael et al., 2005; Deguchi et al., 2005) and multiple sclerosis (MS; Mohan et al., 2010; Sobel, 2001; Sobel and Ahmed, 2001). Deposition of CSPGs post-injury may be a protective CNS response that contains the damage and spares intact tissue from further injury (Galtrey and Fawcett, 2007; Silver and Miller, 2004; Yiu and He, 2006). However, that CSPG response often inhibits both regeneration of axons and remyelination by oligodendrocytes (Davies et al., 1999; Fitch and Silver, 2008; Jones et al., 2003; Lau et al., 2012; Sandvig et al., 2004; Schmalfeldt et al., 2000; Sherman and Back, 2008; Siebert and Osterhout, 2011; Siebert et al., 2011; and others).

In the September issue of Experimental Neurology (2013, vol. 247, pp. 113-121), Pendleton et al. shed light on why remyelination after CNS damage is often incomplete, despite the generation of new oligodendroglia at injury sites. They investigated the effects of several CSPGs (aggrecan, neurocan, and NG2) on oligodendrocyte differentiation, process outgrowth, and myelination. CSPGs inhibited both oligodendrocyte progenitor cells (OPCs) process outgrowth and myelination of DRG neurons in co-culture, without altering OPC differentiation when cultured alone. The inhibition was mediated, at least partially, through the receptor protein tyrosine phosphatase sigma (PTPσ) and the Rho-ROCK pathway. Strategies that target PTPσ to promote remyelination may prove to be beneficial in demyelinating disease such as MS.

CSPGs, the extracellular matrix, and myelination

CSPGs consist of a protein core and a varying number of long sulfated unbranched negatively charged glycosaminoglycan (GAG) chains made up of repeating disaccharide units. For discussions of CSPG structure and function see the many excellent CSPG reviews (Bandtlow and Zimmermann, 2000; Bradbury and Carter, 2011; Busch and Silver, 2007; Fitch and Silver, 2008; Galtrey and Fawcett, 2007; Kwok et al., 2012; Morgenstern et al., 2002; Properzi et al., 2003; Schaefer and Schaefer, 2010; Sharma et al., 2012; Sherman and Back, 2008; Zimmermann and Dours-Zimmermann, 2008). Protein cores with chondroitin sulfate GAGs include the hyalectans (aggrecan, brevican, neurocan, versican), NG2, phosphacan, appican, decorin, biglycan and neuroglycan C. In general, CSPGs have an inhibitory impact on cells (Carbonetto et al., 1983; Iaci et al., 2007; Inatani et al., 2001; Siebert and Osterhout, 2011; Turner et al., 1989 and others; Verna et al., 1989; and others). The sulfation pattern of the GAG chains influences CSPG inhibition, although the core proteins themselves also contribute to proteoglycan function (Castillo et al., 1998; Gama et al., 2006; Garwood et al., 1999; Inatani et al., 2001; Laabs et al., 2007; Nakanishi et al., 2006; Schmalfeldt et al., 2000; Sherman and Back, 2008; Snow et al., 1990). CSPGs bind to cell surface receptors to activate growth-inhibitory pathways, but also interact directly with growth factors, cytokines, and guidance molecules to control their availability to differentiating neurons and glia (Bandtlow and Zimmermann, 2000; Fitch and Silver, 1997; Sanes, 1989; Schaefer and Schaefer, 2010; Schwartz and Domowicz, 2004; Sirko et al., 2010; Yamaguchi, 2000; Zimmermann and Dours-Zimmermann, 2008). Hyalectans, particularly neurocan and brevican, are the predominant CNS CSPGs, while versican and aggrecan are throughout the body. The proteoglycan NG2 is also highly expressed in the CNS, primarily by OPCs and pericytes of the vasculature (Chang et al., 2000; Nishiyama et al., 1999; Ozerdem et al., 2001).

It is difficult to discuss the role of CSPGs in development or after injury without discussing the overall ECM environment. The CNS ECM provides critical structural support, but is also an important signaling scaffold that controls distribution and presentation of soluble molecules in the ECM (Bandtlow and Zimmermann, 2000 and others; Fitch and Silver, 1997; Sanes, 1989; Schaefer and Schaefer, 2010; Schwartz and Domowicz, 2004; Sirko et al., 2010; Yamaguchi, 2000; Zimmermann and Dours-Zimmermann, 2008) and others). It is enriched in proteoglycans and glycoproteins, particularly CSPGs, tenascins, and link proteins that bind hyalectans to hyaluronan, stabilizing ECM structure (Lundell et al., 2004; Yamaguchi, 2000). Collagen I and the small, leucine-rich proteoglycan, decorin, are widespread throughout the ECM, while collagen IV and laminin are expressed in basal laminae (Gordon and Hahn, 2010; Hanemann et al., 1993; Kappler et al., 1998; Villanova et al., 1997; Yurchenco et al., 1992). Fibronectin is expressed in neuroepithelial cells of the ventricular zone and radial glial cells during development, as well as the vasculature (Jones et al., 1982; Pearlman and Sheppard, 1996). During development, the ECM regulates neuronal and glial migration and process outgrowth, with early matrix being loose and allowing extensive cell migration (Carbonetto, 1984; Franco and Müller, 2011; Rutka et al., 1988; Liu et al., 2006; Pearlman and Sheppard, 1996; Sanes, 1989; Schwartz and Domowicz, 2004; and many others). With CNS maturation, the ECM composition changes to form a firmer matrix (Zimmermann and Dours-Zimmermann, 2008).

Besides being critical for CNS architecture and structure, CSPGs play important roles in CNS myelination. In developing white matter tracts, versican and brevican are most abundant (Ogawa et al., 2001; Schmalfeldt et al., 1998), along with the ECM glycoprotein tenascin-R, which is deposited by developing oligodendrocytes (Pesheva and Probstmeier, 2000; Pesheva et al., 1989). During myelination, brevican expression peaks as oligodendrocytes extend membrane processes to ensheath axons (Ogawa et al., 2001). Brevican and tenascin-R are also present at the nodes of Ranvier on large diameter myelinated axons, and the association of tenascin-R with nodes is dependent on brevican expression (Bekku et al., 2009). Thus, node of Ranvier formation is dependent on proper CSPG expression. Oligodendrocytes initially express brevican, but after myelination is complete, brevican is expressed by astrocytes rather than oligodendrocytes in white matter tracts (Ogawa et al., 2001); astrocytes also express numerous other proteins and proteoglycans that influence myelination (Barnett and Linington, 2012; Buttery and ffrench-Constant, 1999; Nash et al., 2010; Watkins et al., 2008; Williams et al., 2007).

CNS injury and CSPG production by astrocytes and oligodendrocytes

After CNS injury, the glial scar contains both reactive astrocytes and recruited OPCs. Astrocytes secrete many ECM molecules during this time, including laminin, fibronectin, and tenascin-C (Liesi et al., 1984; Stoffels et al., 2013; Tang et al., 2003; Zhao et al., 2009). Astrocytes and OPCs both upregulate inhibitory CSPGs, including brevican, neurocan, phosphacan, and versican (Jones et al., 2003; Massey et al., 2008; Rhodes and Fawcett, 2004; Tang et al., 2003), and OPCs also upregulate NG2 (Jones et al., 2002; Rhodes et al., 2006; Tan et al., 2005). Although reactive gliosis becomes a barrier to axon regeneration, it is an important protective response. If dividing astrocytes are ablated after SCI, the blood brain barrier (BBB) fails to reform, neuronal and oligodendroglial cell death increases, the immune response is intensified, the lesion volume is greater, and motor defects worsen (Faulkner et al., 2004). Glial scar formation also initiates revascularization of the injured area, which supplies the surrounding tissue with much needed nutritional, metabolic, and trophic support (Liberto et al., 2004). In contrast to the reactive responses of microglia and astrocytes, OPCs respond to injury with increased NG2 proteoglycan expression only if there is an opening in the BBB, and this response is mediated by entering platelets and macrophages (McTigue et al., 2001; 2006; Rhodes et al., 2006; Tripathi and McTigue, 2007). The role of NG2 expression by OPCs after injury is less clear than that of CSPG expression by astrocytes. NG2 has been reported to inhibit axonal growth (Chen et al., 2002; Jones et al., 2003); however other reports suggest that NG2 may be inhibitory only to specific subpopulations of neurons, and it may actually support neuronal outgrowth in some contexts (Hossain-Ibrahim et al., 2007; Niehaus et al., 1999; Schneider et al., 2001; Yang et al., 2006).

ECM changes and demyelination

In contrast to the dearth of neuronal regeneration after injury, oligodendrocytes have a remarkable ability to regenerate damaged myelin sheaths that surround axons. In MS, repair does occur, and remyelination is seen in shadow plaques (Chang et al., 2012; Raine and Wu, 1993; and others). As the disease progresses, however, remyelination becomes less efficient, resulting in chronic demyelination, progressive neurodegeneration and increased clinical disability (Chang et al., 2002; Franklin, 2002). It is unlikely that this inability to remyelinate results from an absence of OPCs in the adult CNS, as OPCs are present in elderly patients even after decades of disease (Chang et al., 2012; 2002). Thus, it appears that this remyelination failure results from changes in the neurons themselves or in the extracellular environment of the CNS, and understanding these differences may yield valuable insights into the remyelination patterns seen in MS.

ECM changes and demyelination: in vivo animal models

Animal models of demyelination are studied for insight into the demyelination/remyelination processes in MS (Back et al., 2005; Bannerman et al., 2006; Fancy et al., 2011; Franklin et al., 1996; 1991; Harrington et al., 2010; Kawczak et al., 1998; Kotter et al., 2001; Lock et al., 2002; Mann et al., 2012; Nait-Oumesmar et al., 1999; Schonberg et al., 2007; Shields et al., 2000; Skihar et al., 2009; Sloane et al., 2010; Talbott et al., 2005; Ulrich and Bornstein, 1973; Zhao et al., 2009). Toxin-induced demyelination following lysolecithin injection into the spinal cord or corpus callosum, or ethidium bromide injection into the cerebellar peduncles results in minimal immune response and robust remyelination (Back et al., 2005; Franklin et al., 1996; 1991; Harrington et al., 2010; Kotter et al., 2001; Nait-Oumesmar et al., 1999; Schonberg et al., 2007; Shields et al., 2000; Skihar et al., 2009; Sloane et al., 2010; Talbott et al., 2005; Zhao et al., 2009). By contrast, immune-mediated demyelination models including experimental allergic encephalomyelitis (EAE) and Theiler's murine encephalomyelitis (TME) often have insufficient remyelination of demyelinated areas (Bannerman et al., 2006; Haist et al., 2012; Kawczak et al., 1998; Lock et al., 2002; Mann et al., 2012; Massa et al., 2002; Ulrich and Bornstein, 1973). Toxin-induced models create precise focal lesions and are used to study remyelination per se, but immune-mediated models more closely resemble the lesions seen in MS. Although none of the animal models recapitulates MS completely, EAE has been effective at identifying several MS therapies, including glatiramer acetate, mitoxantrone, and natalizumab (for review see Denic et al., 2011). After demyelination in either type of model, the composition of the ECM changes, and each model has important spatiotemporal differences in the specific ECM changes. These differences may underlie the relative rates of remyelination success seen across demyelination models.

After injection of ethidium bromide into the cerebellar peduncles, rapid demyelination occurs, quickly followed by robust remyelination (Zhao et al., 2009). Following demyelination, OPCs upregulate αV integrin and migrate into the lesion. In early lesions, invading macrophages also express αV integrin, although at lower levels (Zhao et al., 2009). Once OPC proliferation is complete and OPCs begin to differentiate, αV integrin expression decreases. During early repair, many ECM proteins are upregulated in the lesion, including those that bind αV integrins, such as osteopontin, vitronectin, and fibronectin, which has the potential to impact OPC responses significantly (Zhao et al., 2009; 2008). Fibronectin leaks in from the vasculature, which may initially be permissive for OPC survival and migration (Hu et al., 2009; Stoffels et al., 2013). In early lesions, upregulation of tenascin-C by astrocytes and of vitronectin by microglia/macrophages may be supportive of OPC migration into lesions (Milner et al., 1996). At later time points, increased expression of tenascin-R and laminin-2 (merosin) correlates with OPC differentiation and axonal contact.

In immune-mediated demyelination models, there are also ECM protein and CSPG changes (Haist et al., 2012; Haylock-Jacobs et al., 2011; Sajad et al., 2011). Normally, meninges and blood vessels express collagen IV, entactin, and laminin, while meningeal connective tissues express collagen I, decorin, and fibronectin. Tenascin-C is expressed diffusely though both the grey and white matter and weakly in the meninges. After demyelination, in contrast to toxin-induced lesions, there is prolonged activation of astrocytes, and the basement membrane molecules collagen IV and laminin deposit perivascularly and accumulate in the lesions (Haist et al., 2012). Non-basement membrane molecules only accumulate in nonvascular patterns in demyelinated areas; tenascin-C is deposited first, followed by collagen, decorin and neurocan (Haist et al., 2012). Phosphacan progressively decreases over the course of TME. Fibronectin only increases in the lesion at late stages and is sometimes seen in the cytoplasm of macrophages

A significant difference between toxin- and immune-induced demyelination is astrocyte ECM deposition. In the absence of inflammation, astrocytes secrete ECM proteins that are supportive to OPC migration, survival, and proliferation. Neurons then respond with increased merosin secretion, which supports continued oligodendrocyte survival, maturation and myelination (Buttery and ffrench-Constant, 1999). In contrast, in the face of prolonged inflammation in EAE and TME, astrocytes change their ECM molecule secretion to create a barrier that could halt the immune assault (Haylock-Jacobs et al., 2011; Zhou et al., 2010). A consequence of this altered ECM deposition and increased CSPG level is inhibition of oligodendrocyte process outgrowth, differentiation, and myelination, which could underlie the decreased remyelination (Lau et al., 2012; Pendleton et al., 2013; Siebert and Osterhout, 2011).

Developing effective remyelination therapies will require understanding astrocyte responses to inflammation in demyelinated lesions. Since current MS therapies dampen the inflammatory response in patients, reducing the immune barrier function of the astrocyte-secreted ECM molecules could be an option. Thus, new repair therapies will require strategies to decrease inhibitory CSPG production by astrocytes and increase the deposition of ECM molecules that promote oligodendrocyte migration, maturation, survival and ultimately the myelination of neurons, as seen in non-immune mediated demyelination paradigms.

ECM changes and demyelination: multiple sclerosis

In MS patients, ECM composition changes occur in both active and chronic lesions, as well as in the normal appearing white matter (Gutowski et al., 1999; Mohan et al., 2010; Sobel, 2001; 1998; Sobel and Ahmed, 2001). ECM molecules that normally contribute to basement membranes, including laminin, collagen type IV, as well as heparan sulfate proteoglycans (HSPGs) and fibronectin, are abnormally expressed and form dense networks in active, but not pre-active or chronic MS lesions (Kurazono et al., 2001; Okamoto et al., 2003; van Horssen et al., 2006). In active MS lesions, both extracellular fibronectin and fibronectin deposited on vessel walls increases and the amount of fibronectin correlates with the degree of inflammation (Sobel and Mitchell, 1989). Macrophages express fibronectin receptors, as well as fibronectin mRNA, so fibronectin may be important for macrophage migration into the CNS. Astrocytes also synthesize fibronectin, which may be part of the gliosis in MS lesions. Fibronectin does not appear to impede remyelination unless it forms aggregates, which occurs in MS lesions but not in toxin-induced demyelination (Stoffels et al., 2013; Zhao et al., 2009). In active, but not chronic, demyelinated MS lesions, dystrophic, demyelinated axons, astrocytes and the microvasculature also express vitronectin (DeWitt et al., 1994; Sobel et al., 1995).

Besides adding to ECM structural stability, tenascins influence oligodendrocyte survival, proliferation, migration, differentiation and morphological maturation (Czopka et al., 2009; Garcion et al., 2001; Garwood et al., 2004; Kiernan et al., 1996). In acute MS lesions, tenascin-C and tenascin-R glycoproteins are lost, and the loss extends into the normal appearing white matter (Gutowski et al., 1999). The loss of tenascin-R and tenascin-C in demyelinated lesions coincides with the greatest density of macrophages (Gutowski et al., 1999). In subacute lesions, tenascin-C and tenascin-R immunoreactive astrocytes are present, particularly in the center of the lesion, while in chronic lesions, tenascin-C and tenascin-R are expressed at levels comparable to the adjacent white matter. Tenascins influence the reactivity of astrocytes, which are a major source of CSPGs after injury (Galtrey and Fawcett, 2007; Holley et al., 2005; Silver and Miller, 2004; Yiu and He, 2006), suggesting that the spatiotemporal expression of tenascin-C and tenascin-R in lesions may be important determinants of remyelination.

CSPGs themselves are also upregulated in demyelinated MS lesions (Haist et al., 2012; Haylock-Jacobs et al., 2011; Lau et al., 2012; Sobel and Ahmed, 2001; van Horssen et al., 2006). Specifically, astrocytes at the edge of active MS lesions upregulate versican, aggrecan, and neurocan. These CSPGs are decreased in the lesion center, and mostly absent from inactive lesions, while phosphacan is unaltered (Crespo et al., 2007; Galtrey and Fawcett, 2007; Rhodes and Fawcett, 2004; Sobel and Ahmed, 2001; van Horssen et al., 2006). It is intriguing to note that several of these same CSPGs, namely neurocan, aggrecan and NG2, impair oligodendrocyte process outgrowth and myelination (Lau et al., 2012; Pendleton et al., 2013; Siebert and Osterhout, 2011).

Aberrant deposition of ECM matrix molecules occurs by leakage from the damaged vasculature and BBB disruption, and by altered secretion from reactive astrocytes. The aberrant deposition of ECM molecules then creates a lesion environment that is inhibitory to oligodendrocyte process extension and remyelination of axons, which in turn leads to increasing neuronal damage as the disease progresses. These are dynamic processes and the level of each ECM molecule changes as lesions progress and recover. The degree of remyelination likely results from the net ratios of supportive (unaggregated fibronectin, merosin) and inhibitory (CSPGs, tenascins, fibronectin aggregates) ECM molecules within and around the lesion.

These ECM changes raise the question of whether ECM deposition impacts leukocyte migration. In active lesions, the integrin receptors for vitronectin (αVβ1) are increased on macrophages and glial cells, as well as on platelets and endothelial cells (Sobel et al., 1995; Zhao et al., 2009; 2008). Vitronectin or fibronectin deposition promotes B cell migration and activation in vitro (Salcedo and Patarroyo, 1995; Shibayama et al., 1995). Thus, ECM changes may allow immune cells to enter and invade lesions, and may even aid early OPC migration into demyelinated areas, while inhibiting later neuronal and oligodendroglia regeneration responses. The data suggest that leukocyte and macrophage invasion results in the destruction of the normal ECM in CNS white and grey matter.

Oligodendrocyte responses to demyelination and CSPGs

After injury, reactive astrocytes and macrophages upregulate CSPGs, and OPCs themselves become reactive, change morphologically, and increase expression of CSPGs, such as NG2 and versican (Chen et al., 2002; Tan et al., 2006). Thus, oligodendrocytes are exposed to very high levels of CSPGs after CNS damage. Following lysolecithin- or SCI-induced demyelination, CSPGs negatively influence oligodendrocyte remyelination of spared axons (Lau et al., 2012; Pendleton et al., 2013; Siebert et al., 2011). Oligodendrocytes share with neurons many of the inhibitory effects, receptors, and downstream signaling events in response to CSPGs. Thus, extensive information can be gleaned from the effects of CSPGs on neurons. Neurons can bind CSPGs via several recently identified receptors, such as the Nogo receptor family members, NgR1 and NgR2, and the leukocyte common antigen-related subfamily of receptor protein tyrosine phosphatases, LAR and PTPσ (Dickendesher et al., 2012; Fisher et al., 2011; Shen et al., 2009). PTPσ was originally identified as a neuronal receptor for heparan sulfate proteoglycans (HSPGs; Aricescu et al., 2002). PTPσ-HSPG binding plays a positive role in axon guidance promoting outgrowth and synapse formation (Aricescu et al., 2002; Coles et al., 2011; Horn et al., 2012), while PTPσ-CSPG binding mediates inhibition of axonal growth and neural regeneration (Coles et al., 2011; Shen et al., 2009). When bound to HSPGs, PTPσ oligomers expressed by neurons sequester in microdomains containing high phosphotyrosine-modified protein, which promotes neuronal outgrowth. On the other hand, when bound to CSPGs, PTPσ is not tightly oligomerized and increased phosphatase activity reduces phosphorylation and inhibits axon growth (Coles et al., 2011). Whether or not PTPσ exhibits contrasting effects on oligodendrocyte process outgrowth and myelination when bound to HSPGs, as seen in neurons, remains to be determined.

Although it has been known that oligodendrocytes express PTPσ (Chesini et al., 2011; Harroch et al., 2002; Kuboyama et al., 2012; Lamprianou et al., 2011; Lamprianou and Harroch, 2006), the role of PTPσ in oligodendrocyte development was not reported until the current study by Pendleton et al. (2013). In this study, they show that CSPGs bound to PTPσ inhibit oligodendroglial process outgrowth. Oligodendrocytes cultured alone on mixtures of laminin and either aggrecan, neurocan, or NG2 had reduced process outgrowth. CSPG digestion with chondroitinase ABC (chABC), which cleaves the GAG side chains from the core protein (Bradbury et al., 2002; Huang et al., 2006), reduced the inhibition on process extension. Knockdown of PTPσ also reversed the inhibition of process outgrowth on CSPGs. PTPσ-CSPG binding also regulated DRG axon myelination in co-culture. Thus, pretreatment of wild type oligodendrocytes with the CSPG aggrecan reduced myelination of DRG axons, but aggrecan did not reduce axon myelination by PTPσ-/- oligodendrocytes. Intriguingly, OPCs cultured on aggregan did not reduce their expression of differentiation markers. Thus, PTPσ activation by CSPGs likely impedes myelination by interfering with process extension towards axons, rather than by inhibiting oligodendrocyte differentiation directly. As PTPσ has also been shown to be an important neuronal receptor for mediating the inhibitory actions of CSPGs on neuronal regeneration (Fry et al., 2010; Shen et al., 2009), these results suggest that strategies targeting PTPσ to promote axonal regeneration may have the added benefit of promoting remyelination. In addition, PTPσ may prove to be a viable new therapeutic target for demyelinating disease such as MS.

OPCs express low levels of PTPs, which increase as they mature into oligodendrocytes (Ranjan and Hudson, 1996). Each PTP has a unique impact on oligodendrocyte development. Early OPCs express PTPα, which is dramatically upregulated as the cells mature. PTPβ/ζ and PTPγ also increase in differentiated oligodendrocytes. Alternative splicing of PTPβ/ζ generates a secreted extracellular domain product, also known as phosphacan (Maurel et al., 1994). Only mature oligodendrocytes express PTPε, which is closely related to PTPα. PTP-specific knockout studies demonstrate that each PTP exerts control over a distinct set of oligodendrocyte activities. PTPα has a negative effect on proliferation and a positive effect on differentiation. Cultures of PTPα-/- OPCs have enhanced proliferation and survival and reduced growth factor dependence. PTPα-/- mice have increased oligodendrocyte lineage cells in the embryonic forebrain and delayed OPC development, with fewer mature oligodendrocytes in the early postnatal corpus callosum (Wang et al., 2012). In contrast, PTPβ/ζ negatively regulates proliferation as well as oligodendrocyte differentiation. PTPβ/ζ-deficient mice have increased proliferation and early differentiation of OPCs in both the brain and optic nerve. At P10, PTPβ/ζ- /- mice have more myelinated axons in the corpus callosum, though normal levels by 3 months of age (Kuboyama et al., 2012; Wang et al., 2012). PTPε, on the other hand, appears to promote oligodendrocyte differentiation and myelination. While most CNS myelin tracts are normal in mice expressing an inactive form of PTPε, there are lower numbers of differentiated oligodendrocytes and delayed myelination of the optic nerve (Muja et al., 2004). However, since transgene expression also occurs in the retinal ganglion cells whose axons were abnormally myelinated, disruption of PTPε signaling occurred in both cell types, complicating interpretation. PTPβ/ζ is important for remyelination by oligodendrocytes following demyelination in EAE. However, there are conflicting reports about whether the suppression of PTPβ/ζ is beneficial and promotes oligodendrocyte survival (Kuboyama et al., 2012), or detrimental and increases apoptosis of mature oligodendrocytes (Harroch et al., 2002). In culture of human cells, PTPβ/ζ suppression maintains the self-renewal capacity of OPCs through activation of the β-catenin/TCF transcription (McClain et al., 2012).

Many of the downstream pathways regulated by PTPs in oligodendrocytes are known to influence oligodendrocyte proliferation, survival, differentiation, and myelination, in particular signaling through Rho, Rac, Cdc42 and the Rho-associated protein kinase (ROCK). For example, major substrates of PTPα are the Src family kinases (Pallen, 2003). In primary OPCs, PTPα acts through the Src kinase Fyn to negatively regulate the activity of Rho, Rac1, and Cdc42, thereby limiting proliferation and promoting differentiation (Pallen, 2003; Wang et al., 2012). In contrast, in differentiating oligodendrocytes, PTPα positively regulates Rac1 and Cdc42, while negatively regulating Rho signaling, which promotes cytoskeletal rearrangements necessary for process extension and maturation (Wang et al., 2012; 2009). Thus, the effect of PTPα on downstream GTPase activity varies in proliferating versus differentiating oligodendroglial cells. These changes in PTP responses could result from extracellular activation of other transmembrane receptors, such as growth factor receptor receptors, or integrins, as PTPs can be activated or inactivated by both of these receptor types (Mattila et al., 2005; Schlessinger et al., 1995; Wang et al., 2009). CSPG-mediated inhibition of oligodendrocyte process outgrowth acts through ROCK, and in vitro studies show that inhibition of ROCK can reverse the inhibitory effects of CSPGs on oligodendrocyte process outgrowth and differentiation (Pendleton et al., 2013; Siebert and Osterhout, 2011). Since OPCs express only low levels of CSPG receptors, they are able to migrate into active lesion sites, despite the high CSPG concentration. As OPCs mature and increase PTPσ expression, they would become more responsive to CSPGs present in the lesion that inhibit oligodendrocyte process extension and maturation. If PTPσ also has distinct roles at different phases of oligodendrocyte development, as the results of Pendleton et al. (2013) suggest, it will be important to determine these differences prior to considering PTPσ in oligodendrocytes as a therapeutic target.

Clinical implications

Multiple studies have demonstrated the benefit of reduced CSPG accumulation and signaling for both axonal regeneration and remyelination (Bradbury et al., 2002; Davies et al., 2004; Huang et al., 2006; Karimi-Abdolrezaee et al., 2012; Siebert et al., 2011; Soleman et al., 2012; Starkey et al., 2012; Tester and Howland, 2008). After injury, chABC-induced removal of the GAG side chains from the core CSPG protein reverses the inhibited axonal regeneration and improves functional outcomes in SCI (Barritt et al., 2006; Bradbury et al., 2002; Huang et al., 2006; Karimi-Abdolrezaee et al., 2012; Siebert et al., 2011; Soleman et al., 2012; Starkey et al., 2012; Tester and Howland, 2008; and others). As noted above, in injury paradigms, CSPGs also inhibit oligodendrocyte differentiation and remyelination. Thus, reducing CSPG deposition can be beneficial for remyelination by enhancing OPC migration into the injury site and promoting endogenous OPC differentiation and maturation (Karimi-Abdolrezaee et al., 2012; Lau et al., 2012; Pendleton et al., 2013; Siebert et al., 2011; Siebert and Osterhout, 2011). Following spinal contusion injury, an immediate local injection of chABC at the site of SCI results in CSPG degradation and increases the number of migratory OPCs near the injury site, particularly at the lesion periphery (Siebert et al., 2011). Intrathecal application of chABC four days after SCI, along with growth factor treatment, increases the proliferation of neural stem cells in the cord, and promotes their differentiation into oligodendrocytes rather than astrocytes (Karimi-Abdolrezaee et al., 2012). In a study that specifically examined remyelination after lysolecithin-induced demyelination, CSPGs accumulated in astrocytes at the lesion edge and microglia in the lesion center at 7 days post injection (DPI, (Lau et al., 2012). At 14 DPI, CSPGs were reduced and they were cleared by the time of remyelination at 21 DPI. If the mice were systemically treated with xyloside, a glycoside that prevents the attachment of GAG side chains on to the CSPG protein cores, the level of CSPGs was reduced and the number of OPCs was increased at 7 DPI. At 21 DPI, the number of mature oligodendrocytes increased and the area of the dorsal columns that became remyelinated also increased.

These studies demonstrate that reducing CSPG levels in demyelinated lesions can promote axonal remyelination. Although removal of CSPGs with chABC has successfully improved recovery in various injury models, it would be difficult to translate this approach to the clinic. The bacterial enzyme requires continuous application or repeated injections into the site of injury to be effective and this could lead to an immune response directed against chABC. In MS, where lesions are diffuse and often hard to identify or access, a systemic treatment such as xyloside could be more beneficial but equally broad-acting. However, the study by Pendleton et al. (2013) suggests that targeting the downstream signaling pathway from the CSPG receptor PTPσ might be a more amenable approach to reducing CSPG inhibition of remyelination. Furthermore, a recent study identified small molecule inhibitors that are selective for PTPσ (Martin et al., 2012), which raises the intriguing possibility that such a tailored therapy may be on the horizon.

Literature Cited

  1. Aricescu AR, McKinnell IW, Halfter W, Stoker AW. Heparan sulfate proteoglycans are ligands for receptor protein tyrosine phosphatase sigma. Mol Cell Biol. 2002;22:1881–1892. doi: 10.1128/MCB.22.6.1881-1892.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Back SA, Tuohy TMF, Chen H, Wallingford N, Craig A, Struve J, Luo NL, Banine F, Liu Y, Chang A, Trapp BD, Bebo BF, Rao MS, Sherman LS. Hyaluronan accumulates in demyelinated lesions and inhibits oligodendrocyte progenitor maturation. Nat Med. 2005;11:966–972. doi: 10.1038/nm1279. [DOI] [PubMed] [Google Scholar]
  3. Bandtlow CE, Zimmermann DR. Proteoglycans in the developing brain: new conceptual insights for old proteins. Physiol Rev. 2000;80:1267–1290. doi: 10.1152/physrev.2000.80.4.1267. [DOI] [PubMed] [Google Scholar]
  4. Bannerman P, Hahn A, Soulika A, Gallo V, Pleasure D. Astrogliosis in EAE spinal cord: Derivation from radial glia, and relationships to oligodendroglia. Glia. 2006;55:57–64. doi: 10.1002/glia.20437. [DOI] [PubMed] [Google Scholar]
  5. Barnett SC, Linington C. Myelination: Do Astrocytes Play a Role? Neuroscientist. 2012 doi: 10.1177/1073858412465655. [DOI] [PubMed] [Google Scholar]
  6. Barritt AW, Davies M, Marchand F, Hartley R, Grist J, Yip P, McMahon SB, Bradbury EJ. Chondroitinase ABC Promotes Sprouting of Intact and Injured Spinal Systems after Spinal Cord Injury. Journal of Neuroscience. 2006;26:10856–10867. doi: 10.1523/JNEUROSCI.2980-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Bekku Y, Rauch U, Ninomiya Y, Oohashi T. Brevican distinctively assembles extracellular components at the large diameter nodes of Ranvier in the CNS. J Neurochem. 2009;108:1266–1276. doi: 10.1111/j.1471-4159.2009.05873.x. [DOI] [PubMed] [Google Scholar]
  8. Bradbury EJ, Carter LM. Manipulating the glial scar: chondroitinase ABC as a therapy for spinal cord injury. Brain Res Bull. 2011;84:306–316. doi: 10.1016/j.brainresbull.2010.06.015. [DOI] [PubMed] [Google Scholar]
  9. Bradbury EJ, Moon LDF, Popat RJ, King VR, Bennett GS, Patel PN, Fawcett JW, McMahon SB. Chondroitinase ABC promotes functional recovery after spinal cord injury. Nature. 2002;416:636–640. doi: 10.1038/416636a. [DOI] [PubMed] [Google Scholar]
  10. Busch SA, Silver J. The role of extracellular matrix in CNS regeneration. Curr Opin Neurobiol. 2007;17:120–127. doi: 10.1016/j.conb.2006.09.004. [DOI] [PubMed] [Google Scholar]
  11. Buttery PC, ffrench-Constant C. Laminin-2/integrin interactions enhance myelin membrane formation by oligodendrocytes. Mol Cell Neurosci. 1999;14:199–212. doi: 10.1006/mcne.1999.0781. [DOI] [PubMed] [Google Scholar]
  12. Carbonetto S. The extracellular matrix of the nervous system. Trends Neurosci. 1984;7:382–387. [Google Scholar]
  13. Carbonetto S, Gruver MM, Turner DC. Nerve fiber growth in culture on fibronectin, collagen, and glycosaminoglycan substrates. J Neurosci. 1983;3:2324–2335. doi: 10.1523/JNEUROSCI.03-11-02324.1983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Carmichael ST, Archibeque I, Luke L, Nolan T, Momiy J, Li S. Growth-associated gene expression after stroke: evidence for a growth-promoting region in peri-infarct cortex. Exp Neurol. 2005;193:291–311. doi: 10.1016/j.expneurol.2005.01.004. [DOI] [PubMed] [Google Scholar]
  15. Castillo GM, Cummings JA, Yang W, Judge ME, Sheardown MJ, Rimvall K, Hansen JB, Snow AD. Sulfate content and specific glycosaminoglycan backbone of perlecan are critical for perlecan's enhancement of islet amyloid polypeptide (amylin) fibril formation. Diabetes. 1998;47:612–620. doi: 10.2337/diabetes.47.4.612. [DOI] [PubMed] [Google Scholar]
  16. Chang A, Nishiyama A, Peterson J, Prineas J, Trapp BD. NG2-Positive Oligodendrocyte Progenitor Cells in Adult Human Brain and Multiple Sclerosis Lesions. The Journal of. 2000;20:6404–6412. doi: 10.1523/JNEUROSCI.20-17-06404.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Chang A, Staugaitis SM, Dutta R, Batt CE, Easley KE, Chomyk AM, Yong VW, Fox RJ, Kidd GJ, Trapp BD. Cortical remyelination: A new target for repair therapies in multiple sclerosis. Ann Neurol. 2012;72:918–926. doi: 10.1002/ana.23693. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Chang A, Tourtellotte WW, Rudick R, Trapp BD. Premyelinating oligodendrocytes in chronic lesions of multiple sclerosis. N Engl J Med. 2002;346:165–173. doi: 10.1056/NEJMoa010994. [DOI] [PubMed] [Google Scholar]
  19. Chen ZJ, Ughrin Y, Levine JM. Inhibition of axon growth by oligodendrocyte precursor cells. Mol Cell Neurosci. 2002;20:125–139. doi: 10.1006/mcne.2002.1102. [DOI] [PubMed] [Google Scholar]
  20. Chesini IM, Debyser G, Croes H, Dam, Ten GB, Devreese B, Stoker AW, Hendriks WJAJ. PTPBR7 binding proteins in myelinating neurons of the mouse brain. Int. J. Biol. Sci. 2011;7:978–991. doi: 10.7150/ijbs.7.978. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Coles CH, Shen Y, Tenney AP, Siebold C, Sutton GC, Lu W, Gallagher JT, Jones EY, Flanagan JG, Aricescu AR. Proteoglycan-specific molecular switch for RPTPσ clustering and neuronal extension. Science. 2011;332:484–488. doi: 10.1126/science.1200840. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Crespo D, Asher RA, Lin R, Rhodes KE, Fawcett JW. How does chondroitinase promote functional recovery in the damaged CNS? Exp Neurol. 2007;206:159–171. doi: 10.1016/j.expneurol.2007.05.001. [DOI] [PubMed] [Google Scholar]
  23. Czopka T, Holst, von A, Schmidt G, ffrench-Constant C, Faissner A. Tenascin C and tenascin R similarly prevent the formation of myelin membranes in a RhoA-dependent manner, but antagonistically regulate the expression of myelin basic protein via a separate pathway. Glia. 2009;57:1790–1801. doi: 10.1002/glia.20891. [DOI] [PubMed] [Google Scholar]
  24. Davies JE, Tang X, Denning JW, Archibald SJ, Davies SJA. Decorin suppresses neurocan, brevican, phosphacan and NG2 expression and promotes axon growth across adult rat spinal cord injuries. Eur J Neurosci. 2004;19:1226–1242. doi: 10.1111/j.1460-9568.2004.03184.x. [DOI] [PubMed] [Google Scholar]
  25. Davies SJA, Goucher DR, Doller C, Silver J. Robust regeneration of adult sensory axons in degenerating white matter of the adult rat spinal cord. Journal of Neuroscience. 1999;19:5810–5822. doi: 10.1523/JNEUROSCI.19-14-05810.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Deguchi K, Takaishi M, Hayashi T, Oohira A, Nagotani S, Li F, Jin G, Nagano I, Shoji M, Miyazaki M, Abe K, Huh N-H. Expression of neurocan after transient middle cerebral artery occlusion in adult rat brain. Brain Res. 2005;1037:194–199. doi: 10.1016/j.brainres.2004.12.016. [DOI] [PubMed] [Google Scholar]
  27. Denic A, Johnson AJ, Bieber AJ, Warrington AE, Rodriguez M, Pirko I. The relevance of animal models in multiple sclerosis research. Pathophysiology. 2011;18:21–29. doi: 10.1016/j.pathophys.2010.04.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. DeWitt DA, Richey PL, Praprotnik D, Silver J, Perry G. Chondroitin sulfate proteoglycans are a common component of neuronal inclusions and astrocytic reaction in neurodegenerative diseases. Brain Res. 1994;656:205–209. doi: 10.1016/0006-8993(94)91386-2. [DOI] [PubMed] [Google Scholar]
  29. DeWitt DA, Silver J, Canning DR, Perry G. Chondroitin sulfate proteoglycans are associated with the lesions of Alzheimer's disease. Exp Neurol. 1993;121:149–152. doi: 10.1006/exnr.1993.1081. [DOI] [PubMed] [Google Scholar]
  30. Dickendesher TL, Baldwin KT, Mironova YA, Koriyama Y, Raiker SJ, Askew KL, Wood A, Geoffroy CG, Zheng B, Liepmann CD, Katagiri Y, Benowitz LI, Geller HM, Giger RJ. NgR1 and NgR3 are receptors for chondroitin sulfate proteoglycans. Nature Publishing Group. 2012;15:703–712. doi: 10.1038/nn.3070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Fancy SPJ, Harrington EP, Yuen TJ, Silbereis JC, Zhao C, Baranzini SE, Bruce CC, Otero JJ, Huang EJ, Nusse R, Franklin RJM, Rowitch DH. Axin2 as regulatory and therapeutic target in newborn brain injury and remyelination. Nature Neuroscience. 2011;14:1009–1016. doi: 10.1038/nn.2855. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Faulkner JR, Herrmann JE, Woo MJ, Tansey KE, Doan NB, Sofroniew MV. Reactive astrocytes protect tissue and preserve function after spinal cord injury. Journal of Neuroscience. 2004;24:2143–2155. doi: 10.1523/JNEUROSCI.3547-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Fisher D, Xing B, Dill J, Li H, Hoang HH, Zhao Z, Yang X-L, Bachoo R, Cannon S, Longo FM, Sheng M, Silver J, Li S. Leukocyte common antigen-related phosphatase is a functional receptor for chondroitin sulfate proteoglycan axon growth inhibitors. Journal of Neuroscience. 2011;31:14051–14066. doi: 10.1523/JNEUROSCI.1737-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Fitch MT, Silver J. Glial cell extracellular matrix: boundaries for axon growth in development and regeneration. Cell Tissue Res. 1997;290:379–384. doi: 10.1007/s004410050944. [DOI] [PubMed] [Google Scholar]
  35. Fitch MT, Silver J. CNS injury, glial scars, and inflammation: Inhibitory extracellular matrices and regeneration failure. Exp Neurol. 2008;209:294–301. doi: 10.1016/j.expneurol.2007.05.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Franco SJ, Müller U. Extracellular matrix functions during neuronal migration and lamination in the mammalian central nervous system. Dev Neurobiol. 2011;71:889–900. doi: 10.1002/dneu.20946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Franklin RJM. Why does remyelination fail in multiple sclerosis? Nat Rev Neurosci. 2002;3:705–714. doi: 10.1038/nrn917. [DOI] [PubMed] [Google Scholar]
  38. Franklin RJM, Bayley SA, Blakemore WF. Transplanted CG4 cells (an oligodendrocyte progenitor cell line) survive, migrate, and contribute to repair of areas of demyelination in X-irradiated and damaged spinal cord but not in normal spinal cord. Exp Neurol. 1996;137:263–276. doi: 10.1006/exnr.1996.0025. [DOI] [PubMed] [Google Scholar]
  39. Franklin RJM, Crang AJ, Blakemore WF. Transplanted type-1 astrocytes facilitate repair of demyelinating lesions by host oligodendrocytes in adult rat spinal cord. J Neurocytol. 1991;20:420–430. doi: 10.1007/BF01355538. [DOI] [PubMed] [Google Scholar]
  40. Fry EJ, Chagnon MJ, López-Vales R, Tremblay ML, David S. Corticospinal tract regeneration after spinal cord injury in receptor protein tyrosine phosphatase sigma deficient mice. Glia. 2010;58:423–433. doi: 10.1002/glia.20934. [DOI] [PubMed] [Google Scholar]
  41. Galtrey CM, Fawcett JW. The role of chondroitin sulfate proteoglycans in regeneration and plasticity in the central nervous system. Brain research reviews. 2007;54:1–18. doi: 10.1016/j.brainresrev.2006.09.006. [DOI] [PubMed] [Google Scholar]
  42. Galtrey CM, Kwok JCF, Carulli D, Rhodes KE, Fawcett JW. Distribution and synthesis of extracellular matrix proteoglycans, hyaluronan, link proteins and tenascin-R in the rat spinal cord. Eur J Neurosci. 2008;27:1373–1390. doi: 10.1111/j.1460-9568.2008.06108.x. [DOI] [PubMed] [Google Scholar]
  43. Gama CI, Tully SE, Sotogaku N, Clark PM, Rawat M, Vaidehi N, Goddard WA, Nishi A, Hsieh-Wilson LC. Sulfation patterns of glycosaminoglycans encode molecular recognition and activity. Nat Chem Biol. 2006;2:467–473. doi: 10.1038/nchembio810. [DOI] [PubMed] [Google Scholar]
  44. Garcion E, Faissner A, ffrench-Constant C. Knockout mice reveal a contribution of the extracellular matrix molecule tenascin-C to neural precursor proliferation and migration. Development. 2001;128:2485–2496. doi: 10.1242/dev.128.13.2485. [DOI] [PubMed] [Google Scholar]
  45. Garwood J, Garcion E, Dobbertin A, Heck N, Calco V, ffrench-Constant C, Faissner A. The extracellular matrix glycoprotein Tenascin-C is expressed by oligodendrocyte precursor cells and required for the regulation of maturation rate, survival and responsiveness to platelet-derived growth factor. Eur J Neurosci. 2004;20:2524–2540. doi: 10.1111/j.1460-9568.2004.03727.x. [DOI] [PubMed] [Google Scholar]
  46. Garwood J, Schnädelbach O, Clement A, Schütte K, Bach A, Faissner A. DSD-1-proteoglycan is the mouse homolog of phosphacan and displays opposing effects on neurite outgrowth dependent on neuronal lineage. Journal of Neuroscience. 1999;19:3888–3899. doi: 10.1523/JNEUROSCI.19-10-03888.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Gordon MK, Hahn RA. Collagens. Cell Tissue Res. 2010;339:247–257. doi: 10.1007/s00441-009-0844-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Gutowski NJ, Newcombe J, Cuzner ML. Tenascin-R and C in multiple sclerosis lesions: relevance to extracellular matrix remodelling. Neuropathol. Appl. Neurobiol. 1999;25:207–214. doi: 10.1046/j.1365-2990.1999.00176.x. [DOI] [PubMed] [Google Scholar]
  49. Haist V, Ulrich R, Kalkuhl A, Deschl U, Baumgärtner W. Distinct spatio-temporal extracellular matrix accumulation within demyelinated spinal cord lesions in Theiler's murine encephalomyelitis. Brain Pathology. 2012;22:188–204. doi: 10.1111/j.1750-3639.2011.00518.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Hanemann CO, Kuhn G, Lie A, Gillen C, Bosse F, Spreyer P, Müller HW. Expression of decorin mRNA in the nervous system of rat. J. Histochem. Cytochem. 1993;41:1383–1391. doi: 10.1177/41.9.8354878. [DOI] [PubMed] [Google Scholar]
  51. Harrington EP, Zhao C, Fancy SPJ, Kaing S, Franklin RJM, Rowitch DH. Oligodendrocyte PTEN is required for myelin and axonal integrity, not remyelination. Ann Neurol. 2010;68:703–716. doi: 10.1002/ana.22090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Harroch S, Furtado GC, Brueck W, Rosenbluth J, Lafaille J, Chao M, Buxbaum JD, Schlessinger J. A critical role for the protein tyrosine phosphatase receptor type Z in functional recovery from demyelinating lesions. Nat Genet. 2002;32:411–414. doi: 10.1038/ng1004. [DOI] [PubMed] [Google Scholar]
  53. Haylock-Jacobs S, Keough MB, Lau L, Yong VW. Chondroitin sulphate proteoglycans: extracellular matrix proteins that regulate immunity of the central nervous system. Autoimmunity Reviews. 2011;10:766–772. doi: 10.1016/j.autrev.2011.05.019. [DOI] [PubMed] [Google Scholar]
  54. Holley JE, Gveric D, Whatmore JL, Gutowski NJ. Tenascin C induces a quiescent phenotype in cultured adult human astrocytes. Glia. 2005;52:53–58. doi: 10.1002/glia.20231. [DOI] [PubMed] [Google Scholar]
  55. Horn KE, Xu B, Gobert D, Hamam BN, Thompson KM, Wu C-L, Bouchard J-F, Uetani N, Racine RJ, Tremblay ML, Ruthazer ES, Chapman CA, Kennedy TE. Receptor protein tyrosine phosphatase sigma regulates synapse structure, function and plasticity. J Neurochem. 2012;122:147–161. doi: 10.1111/j.1471-4159.2012.07762.x. [DOI] [PubMed] [Google Scholar]
  56. Hossain-Ibrahim MK, Rezajooi K, Stallcup WB, Lieberman AR, Anderson PN. Analysis of axonal regeneration in the central and peripheral nervous systems of the NG2-deficient mouse. BMC neuroscience. 2007;8:80. doi: 10.1186/1471-2202-8-80. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Hu J, Deng L, Wang X, Xu XM. Effects of extracellular matrix molecules on the growth properties of oligodendrocyte progenitor cells in vitro. J. Neurosci. Res. 2009;87:2854–2862. doi: 10.1002/jnr.22111. [DOI] [PubMed] [Google Scholar]
  58. Huang W-C, Kuo W-C, Cherng J-H, Hsu S-H, Chen P-R, Huang S-H, Huang M-C, Liu J-C, Cheng H. Chondroitinase ABC promotes axonal re-growth and behavior recovery in spinal cord injury. Biochem Biophys Res Commun. 2006;349:963–968. doi: 10.1016/j.bbrc.2006.08.136. [DOI] [PubMed] [Google Scholar]
  59. Iaci JF, Vecchione AM, Zimber MP, Caggiano AO. Chondroitin sulfate proteoglycans in spinal cord contusion injury and the effects of chondroitinase treatment. J Neurotrauma. 2007;24:1743–1759. doi: 10.1089/neu.2007.0366. [DOI] [PubMed] [Google Scholar]
  60. Inatani M, Honjo M, Otori Y, Oohira A, Kido N, Tano Y, Honda Y, Tanihara H. Inhibitory effects of neurocan and phosphacan on neurite outgrowth from retinal ganglion cells in culture. Invest Ophthalmol Vis Sci. 2001;42:1930–1938. [PubMed] [Google Scholar]
  61. Rutka James T, Apodaca Gerard, Stern Robert, Rosenblum Mark. The extracellular matrix of the central and peripheral nervous systems: structure and function. 1988 doi: 10.3171/jns.1988.69.2.0155. http://dx.doi.org/10.3171/jns.1988.69.2.0155. [DOI] [PubMed]
  62. Jones LL, Sajed D, Tuszynski MH. Axonal regeneration through regions of chondroitin sulfate proteoglycan deposition after spinal cord injury: a balance of permissiveness and inhibition. Journal of Neuroscience. 2003;23:9276–9288. doi: 10.1523/JNEUROSCI.23-28-09276.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Jones LL, Yamaguchi Y, Stallcup WB, Tuszynski MH. NG2 is a major chondroitin sulfate proteoglycan produced after spinal cord injury and is expressed by macrophages and oligodendrocyte progenitors. Journal of Neuroscience. 2002;22:2792–2803. doi: 10.1523/JNEUROSCI.22-07-02792.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Jones TR, Ruoslahti E, Schold SC, Bigner DD. Fibronectin and glial fibrillary acidic protein expression in normal human brain and anaplastic human gliomas. Cancer Res. 1982;42:168–177. [PubMed] [Google Scholar]
  65. Kappler J, Stichel CC, Gleichmann M, Gillen C, Junghans U, Kresse H, Müller HW. Developmental regulation of decorin expression in postnatal rat brain. Brain Res. 1998;793:328–332. doi: 10.1016/s0006-8993(98)00260-1. [DOI] [PubMed] [Google Scholar]
  66. Karimi-Abdolrezaee S, Schut D, Wang J, Fehlings MG. Chondroitinase and growth factors enhance activation and oligodendrocyte differentiation of endogenous neural precursor cells after spinal cord injury. PLoS ONE. 2012;7:e37589. doi: 10.1371/journal.pone.0037589. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Kawczak JA, Mathisen PM, Drazba JA, Fuss B, Macklin WB, Tuohy VK. Digitized image analysis reveals diffuse abnormalities in normal-appearing white matter during acute experimental autoimmune encephalomyelitis. J. Neurosci. Res. 1998;54:364–372. doi: 10.1002/(SICI)1097-4547(19981101)54:3<364::AID-JNR7>3.0.CO;2-#. [DOI] [PubMed] [Google Scholar]
  68. Kiernan BW, Götz B, Faissner A, ffrench-Constant C. Tenascin-C inhibits oligodendrocyte precursor cell migration by both adhesion-dependent and adhesion-independent mechanisms. Mol Cell Neurosci. 1996;7:322–335. doi: 10.1006/mcne.1996.0024. [DOI] [PubMed] [Google Scholar]
  69. Kotter MR, Setzu A, Sim FJ, Van Rooijen N, Franklin RJ. Macrophage depletion impairs oligodendrocyte remyelination following lysolecithin-induced demyelination. Glia. 2001;35:204–212. doi: 10.1002/glia.1085. [DOI] [PubMed] [Google Scholar]
  70. Kuboyama K, Fujikawa A, Masumura M, Suzuki R, Matsumoto M, Noda M. Protein tyrosine phosphatase receptor type z negatively regulates oligodendrocyte differentiation and myelination. PLoS ONE. 2012;7:e48797. doi: 10.1371/journal.pone.0048797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Kurazono S, Okamoto M, Sakiyama J, Mori S, Nakata Y, Fukuoka J, Amano S, Oohira A, Matsui H. Expression of brain specific chondroitin sulfate proteoglycans, neurocan and phosphacan, in the developing and adult hippocampus of Ihara's epileptic rats. Brain Res. 2001;898:36–48. doi: 10.1016/s0006-8993(01)02128-x. [DOI] [PubMed] [Google Scholar]
  72. Kwok JCF, Warren P, Fawcett JW. Chondroitin sulfate: a key molecule in the brain matrix. Int J Biochem Cell Biol. 2012;44:582–586. doi: 10.1016/j.biocel.2012.01.004. [DOI] [PubMed] [Google Scholar]
  73. Laabs TL, Wang H, Katagiri Y, McCann T, Fawcett JW, Geller HM. Inhibiting glycosaminoglycan chain polymerization decreases the inhibitory activity of astrocyte-derived chondroitin sulfate proteoglycans. Journal of Neuroscience. 2007;27:14494–14501. doi: 10.1523/JNEUROSCI.2807-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Lamprianou S, Chatzopoulou E, Thomas J-L, Bouyain S, Harroch S. A complex between contactin-1 and the protein tyrosine phosphatase PTPRZ controls the development of oligodendrocyte precursor cells. Proceedings of the National Academy of Sciences. 2011;108:17498–17503. doi: 10.1073/pnas.1108774108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Lamprianou S, Harroch S. Receptor protein tyrosine phosphatase from stem cells to mature glial cells of the central nervous system. JMN. 2006;29:241–255. doi: 10.1385/JMN:29:3:241. [DOI] [PubMed] [Google Scholar]
  76. Lau LW, Keough MB, Haylock-Jacobs S, Cua R, Döring A, Sloka S, Stirling DP, Rivest S, Yong VW. Chondroitin sulfate proteoglycans in demyelinated lesions impair remyelination. Ann Neurol. 2012;72:419–432. doi: 10.1002/ana.23599. [DOI] [PubMed] [Google Scholar]
  77. Lemons ML, Howland DR, Anderson DK. Chondroitin sulfate proteoglycan immunoreactivity increases following spinal cord injury and transplantation. Exp Neurol. 1999;160:51–65. doi: 10.1006/exnr.1999.7184. [DOI] [PubMed] [Google Scholar]
  78. Liberto CM, Albrecht PJ, Herx LM, Yong VW, Levison SW. Pro-regenerative properties of cytokine-activated astrocytes. J Neurochem. 2004;89:1092–1100. doi: 10.1111/j.1471-4159.2004.02420.x. [DOI] [PubMed] [Google Scholar]
  79. Liesi P, Kaakkola S, Dahl D, Vaheri A. Laminin is induced in astrocytes of adult brain by injury. EMBO J. 1984;3:683–686. doi: 10.1002/j.1460-2075.1984.tb01867.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Liu BP, Cafferty WBJ, Budel SO, Strittmatter SM. Extracellular regulators of axonal growth in the adult central nervous system. Philosophical Transactions of the Royal Society B: Biological Sciences. 2006;361:1593–1610. doi: 10.1098/rstb.2006.1891. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Lock C, Hermans G, Pedotti R, Brendolan A, Schadt E, Garren H, Langer-Gould A, Strober S, Cannella B, Allard J, Klonowski P, Austin A, Lad N, Kaminski N, Galli SJ, Oksenberg JR, Raine CS, Heller R, Steinman L. Gene-microarray analysis of multiple sclerosis lesions yields new targets validated in autoimmune encephalomyelitis. Nat Med. 2002;8:500–508. doi: 10.1038/nm0502-500. [DOI] [PubMed] [Google Scholar]
  82. Lundell A, Olin AI, Mörgelin M, al-Karadaghi S, Aspberg A, Logan DT. Structural basis for interactions between tenascins and lectican C-type lectin domains: evidence for a crosslinking role for tenascins. Structure. 2004;12:1495–1506. doi: 10.1016/j.str.2004.05.021. [DOI] [PubMed] [Google Scholar]
  83. Mann MK, Ray A, Basu S, Karp CL, Dittel BN. Pathogenic and regulatory roles for B cells in experimental autoimmune encephalomyelitis. Autoimmunity. 2012;45:388–399. doi: 10.3109/08916934.2012.665523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Martin KR, Narang P, Xu Y, Kauffman AL, Petit J, Xu HE, Meurice N, MacKeigan JP. Identification of small molecule inhibitors of PTPσ through an integrative virtual and biochemical approach. PLoS ONE. 2012;7:e50217. doi: 10.1371/journal.pone.0050217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Massa PT, Ropka SL, Saha S, Fecenko KL, Beuler KL. Critical role for protein tyrosine phosphatase SHP-1 in controlling infection of central nervous system glia and demyelination by Theiler's murine encephalomyelitis virus. J Virol. 2002;76:8335–8346. doi: 10.1128/JVI.76.16.8335-8346.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Massey JM, Amps J, Viapiano MS, Matthews RT, Wagoner MR, Whitaker CM, Alilain W, Yonkof AL, Khalyfa A, Cooper NGF, Silver J, Onifer SM. Increased chondroitin sulfate proteoglycan expression in denervated brainstem targets following spinal cord injury creates a barrier to axonal regeneration overcome by chondroitinase ABC and neurotrophin-3. Exp Neurol. 2008;209:426–445. doi: 10.1016/j.expneurol.2007.03.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Mattila E, Pellinen T, Nevo J, Vuoriluoto K, Arjonen A, Ivaska J. Negative regulation of EGFR signalling through integrin-alpha1beta1-mediated activation of protein tyrosine phosphatase TCPTP. Nat Cell Biol. 2005;7:78–85. doi: 10.1038/ncb1209. [DOI] [PubMed] [Google Scholar]
  88. Maurel P, Rauch U, Flad M, Margolis RK, Margolis RU. Phosphacan, a chondroitin sulfate proteoglycan of brain that interacts with neurons and neural cell-adhesion molecules, is an extracellular variant of a receptor-type protein tyrosine phosphatase. Proc Natl Acad Sci USA. 1994;91:2512–2516. doi: 10.1073/pnas.91.7.2512. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. McClain CR, Sim FJ, Goldman SA. Pleiotrophin suppression of receptor protein tyrosine phosphatase-β/ζ maintains the self-renewal competence of fetal human oligodendrocyte progenitor cells. Journal of Neuroscience. 2012;32:15066–15075. doi: 10.1523/JNEUROSCI.1320-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. McTigue DM, Tripathi R, Wei P. NG2 colocalizes with axons and is expressed by a mixed cell population in spinal cord lesions. J Neuropathol Exp Neurol. 2006;65:406–420. doi: 10.1097/01.jnen.0000218447.32320.52. [DOI] [PubMed] [Google Scholar]
  91. McTigue DM, Wei P, Stokes BT. Proliferation of NG2-positive cells and altered oligodendrocyte numbers in the contused rat spinal cord. Journal of Neuroscience. 2001;21:3392–3400. doi: 10.1523/JNEUROSCI.21-10-03392.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Milner R, Edwards G, Streuli C, ffrench-Constant C. A role in migration for the alpha V beta 1 integrin expressed on oligodendrocyte precursors. J Neurosci. 1996;16:7240–7252. doi: 10.1523/JNEUROSCI.16-22-07240.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Mohan H, Krumbholz M, Sharma R. Extracellular Matrix in Multiple Sclerosis Lesions: Fibrillar Collagens, Biglycan and Decorin are Upregulated and Associated with Infiltrating Immune Cells - Mohan -2010 - Brain Pathology - Wiley Online Library. Brain. 2010 doi: 10.1111/j.1750-3639.2010.00399.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Morgenstern DA, Asher RA, Fawcett JW. Chondroitin sulphate proteoglycans in the CNS injury response. Prog Brain Res. 2002;137:313–332. doi: 10.1016/s0079-6123(02)37024-9. [DOI] [PubMed] [Google Scholar]
  95. Muja N, Lovas G, Romm E, Machleder D, Ranjan M, Gallo V, Hudson LD. Expression of a catalytically inactive transmembrane protein tyrosine phosphatase epsilon (tm-PTP epsilon) delays optic nerve myelination. Glia. 2004;48:278–297. doi: 10.1002/glia.20078. [DOI] [PubMed] [Google Scholar]
  96. Nait-Oumesmar B, Decker L, Lachapelle F, Avellana-Adalid V, Bachelin C, Van Evercooren AB. Progenitor cells of the adult mouse subventricular zone proliferate, migrate and differentiate into oligodendrocytes after demyelination. Eur J Neurosci. 1999;11:4357–4366. doi: 10.1046/j.1460-9568.1999.00873.x. [DOI] [PubMed] [Google Scholar]
  97. Nakanishi K, Aono S, Hirano K, Kuroda Y, Ida M, Tokita Y, Matsui F, Oohira A. Identification of neurite outgrowth-promoting domains of neuroglycan C, a brain-specific chondroitin sulfate proteoglycan, and involvement of phosphatidylinositol 3-kinase and protein kinase C signaling pathways in neuritogenesis. J Biol Chem. 2006;281:24970–24978. doi: 10.1074/jbc.M601498200. [DOI] [PubMed] [Google Scholar]
  98. Nash B, Ioannidou K, Barnett SC. Astrocyte phenotypes and their relationship to myelination. J Anat. 2010;219:44–52. doi: 10.1111/j.1469-7580.2010.01330.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Niehaus A, Stegmüller J, Diers-Fenger M, Trotter J. Cell-surface glycoprotein of oligodendrocyte progenitors involved in migration. J Neurosci. 1999;19:4948–4961. doi: 10.1523/JNEUROSCI.19-12-04948.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Nishiyama A, Chang A, Trapp BD. NG2+ glial cells: a novel glial cell population in the adult brain. J Neuropathol Exp Neurol. 1999;58:1113–1124. doi: 10.1097/00005072-199911000-00001. [DOI] [PubMed] [Google Scholar]
  101. Ogawa T, Hagihara K, Suzuki M, Yamaguchi Y. Brevican in the developing hippocampal fimbria: differential expression in myelinating oligodendrocytes and adult astrocytes suggests a dual role for brevican in central nervous system fiber tract development. J Comp Neurol. 2001;432:285–295. doi: 10.1002/cne.1103. [DOI] [PubMed] [Google Scholar]
  102. Okamoto M, Sakiyama J, Mori S, Kurazono S, Usui S, Hasegawa M, Oohira A. Kainic acid-induced convulsions cause prolonged changes in the chondroitin sulfate proteoglycans neurocan and phosphacan in the limbic structures. Exp Neurol. 2003;184:179–195. doi: 10.1016/s0014-4886(03)00251-6. [DOI] [PubMed] [Google Scholar]
  103. Ozerdem U, Grako KA, Dahlin-Huppe K, Monosov E, Stallcup WB. NG2 proteoglycan is expressed exclusively by mural cells during vascular morphogenesis. Dev. Dyn. 2001;222:218–227. doi: 10.1002/dvdy.1200. [DOI] [PubMed] [Google Scholar]
  104. Pallen C. Protein Tyrosine Phosphatase simga; (PTPsigma: A Src Family Kinase Activator and Mediator of Multiple Biological Effects. CTMC. 2003;3:821–835. doi: 10.2174/1568026033452320. [DOI] [PubMed] [Google Scholar]
  105. Pearlman AL, Sheppard AM. Extracellular matrix in early cortical development. Prog Brain Res. 1996;108:117–134. [PubMed] [Google Scholar]
  106. Pendleton JC, Shamblott MJ, Gary DS, Belegu V, Hurtado A, Malone ML, McDonald JW. Chondroitin sulfate proteoglycans inhibit oligodendrocyte myelination through PTPσ. Exp Neurol. 2013;247C:113–121. doi: 10.1016/j.expneurol.2013.04.003. [DOI] [PubMed] [Google Scholar]
  107. Pesheva P, Probstmeier R. The yin and yang of tenascin-R in CNS development and pathology. Prog Neurobiol. 2000;61:465–493. doi: 10.1016/s0301-0082(99)00061-1. [DOI] [PubMed] [Google Scholar]
  108. Pesheva P, Spiess E, Schachner M. J1-160 and J1-180 are oligodendrocyte-secreted nonpermissive substrates for cell adhesion. J Cell Biol. 1989;109:1765–1778. doi: 10.1083/jcb.109.4.1765. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Properzi F, Asher RA, Fawcett JW. Chondroitin sulphate proteoglycans in the central nervous system: changes and synthesis after injury. Biochem. Soc. Trans. 2003;31:335–336. doi: 10.1042/bst0310335. [DOI] [PubMed] [Google Scholar]
  110. Raine CS, Wu E. Multiple sclerosis: remyelination in acute lesions. J Neuropathol Exp Neurol. 1993;52:199–204. [PubMed] [Google Scholar]
  111. Ranjan M, Hudson LD. Regulation of tyrosine phosphorylation and protein tyrosine phosphatases during oligodendrocyte differentiation. Mol Cell Neurosci. 1996;7:404–418. doi: 10.1006/mcne.1996.0029. [DOI] [PubMed] [Google Scholar]
  112. Rhodes KE, Fawcett JW. Chondroitin sulphate proteoglycans: preventing plasticity or protecting the CNS? J Anat. 2004;204:33–48. doi: 10.1111/j.1469-7580.2004.00261.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Rhodes KE, Raivich G, Fawcett JW. The injury response of oligodendrocyte precursor cells is induced by platelets, macrophages and inflammation-associated cytokines. NSC. 2006;140:87–100. doi: 10.1016/j.neuroscience.2006.01.055. [DOI] [PubMed] [Google Scholar]
  114. Sajad M, Zargan J, Chawla R, Umar S, Khan HA. Upregulation of CSPG3 accompanies neuronal progenitor proliferation and migration in EAE. J Mol Neurosci. 2011;43:531–540. doi: 10.1007/s12031-010-9476-0. [DOI] [PubMed] [Google Scholar]
  115. Salcedo R, Patarroyo M. Constitutive alpha V beta 3 integrin-mediated adhesion of human lymphoid B cells to vitronectin substrate. Cell. Immunol. 1995;160:165–172. doi: 10.1016/0008-8749(95)80023-c. [DOI] [PubMed] [Google Scholar]
  116. Sandvig A, Berry M, Barrett LB, Butt A, Logan A. Myelin-, reactive glia-, and scar-derived CNS axon growth inhibitors: expression, receptor signaling, and correlation with axon regeneration. Glia. 2004;46:225–251. doi: 10.1002/glia.10315. [DOI] [PubMed] [Google Scholar]
  117. Sanes JR. Extracellular Matrix Molecules that Influence Neural Development. Annu Rev Neurosci. 1989;12:491–516. doi: 10.1146/annurev.ne.12.030189.002423. [DOI] [PubMed] [Google Scholar]
  118. Schaefer L, Schaefer RM. Proteoglycans: from structural compounds to signaling molecules. Cell Tissue Res. 2010;339:237–246. doi: 10.1007/s00441-009-0821-y. [DOI] [PubMed] [Google Scholar]
  119. Schlessinger J, Lax I, Lemmon M. Regulation of growth factor activation by proteoglycans: what is the role of the low affinity receptors? Cell. 1995;83:357–360. doi: 10.1016/0092-8674(95)90112-4. [DOI] [PubMed] [Google Scholar]
  120. Schmalfeldt M, Bandtlow CE, Dours-Zimmermann MT, Winterhalter KH, Zimmermann DR. Brain derived versican V2 is a potent inhibitor of axonal growth. Journal of Cell Science. 2000;113(Pt 5):807–816. doi: 10.1242/jcs.113.5.807. [DOI] [PubMed] [Google Scholar]
  121. Schmalfeldt M, Dours-Zimmermann MT, Winterhalter KH, Zimmermann DR. Versican V2 is a major extracellular matrix component of the mature bovine brain. J Biol Chem. 1998;273:15758–15764. doi: 10.1074/jbc.273.25.15758. [DOI] [PubMed] [Google Scholar]
  122. Schneider S, Bosse F, D'Urso D, Muller H, Sereda MW, Nave K-AA, Niehaus A, Kempf T, Schnolzer M, Trotter J. The AN2 protein is a novel marker for the Schwann cell lineage expressed by immature and nonmyelinating Schwann cells. Journal of Neuroscience. 2001;21:920–933. doi: 10.1523/JNEUROSCI.21-03-00920.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Schonberg DL, Popovich PG, McTigue DM. Oligodendrocyte generation is differentially influenced by toll-like receptor (TLR) 2 and TLR4-mediated intraspinal macrophage activation. J Neuropathol Exp Neurol. 2007;66:1124–1135. doi: 10.1097/nen.0b013e31815c2530. [DOI] [PubMed] [Google Scholar]
  124. Schwartz NB, Domowicz M. Proteoglycans in brain development. Glycoconj J. 2004;21:329–341. doi: 10.1023/B:GLYC.0000046278.34016.36. [DOI] [PubMed] [Google Scholar]
  125. Sharma K, Selzer ME, Li S. Scar-mediated inhibition and CSPG receptors in the CNS. Exp Neurol. 2012;237:370–378. doi: 10.1016/j.expneurol.2012.07.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Shen Y, Tenney AP, Busch SA, Horn KP, Cuascut FX, Liu K, He Z, Silver J, Flanagan JG. PTPsigma is a receptor for chondroitin sulfate proteoglycan, an inhibitor of neural regeneration. Science. 2009;326:592–596. doi: 10.1126/science.1178310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Sherman LS, Back SA. A “GAG” reflex prevents repair of the damaged CNS. Trends Neurosci. 2008;31:44–52. doi: 10.1016/j.tins.2007.11.001. [DOI] [PubMed] [Google Scholar]
  128. Shibayama H, Tagawa S, Hattori H, Inoue R, Katagiri S, Kitani T. Laminin and fibronectin promote the chemotaxis of human malignant plasma cell lines. Blood. 1995;86:719–725. [PubMed] [Google Scholar]
  129. Shields SA, Blakemore WF, Franklin RJM. Schwann cell remyelination is restricted to astrocyte-deficient areas after transplantation into demyelinated adult rat brain. J. Neurosci. Res. 2000;60:571–578. doi: 10.1002/(SICI)1097-4547(20000601)60:5<571::AID-JNR1>3.0.CO;2-Q. [DOI] [PubMed] [Google Scholar]
  130. Siebert JR, Osterhout DJ. The inhibitory effects of chondroitin sulfate proteoglycans on oligodendrocytes. J Neurochem. 2011;119:176–188. doi: 10.1111/j.1471-4159.2011.07370.x. [DOI] [PubMed] [Google Scholar]
  131. Siebert JR, Stelzner DJ, Osterhout DJ. Chondroitinase treatment following spinal contusion injury increases migration of oligodendrocyte progenitor cells. Exp Neurol. 2011;231:19–29. doi: 10.1016/j.expneurol.2011.05.002. [DOI] [PubMed] [Google Scholar]
  132. Silver J, Miller JH. Regeneration beyond the glial scar. Nat Rev Neurosci. 2004;5:146–156. doi: 10.1038/nrn1326. [DOI] [PubMed] [Google Scholar]
  133. Sirko S, Holst, von A, Weber A, Wizenmann A, Theocharidis U, Götz M, Faissner A. Chondroitin sulfates are required for fibroblast growth factor-2-dependent proliferation and maintenance in neural stem cells and for epidermal growth factor-dependent migration of their progeny. Stem Cells. 2010;28:775–787. doi: 10.1002/stem.309. [DOI] [PubMed] [Google Scholar]
  134. Skihar V, Silva C, Chojnacki A, Döring A, Stallcup WB, Weiss S, Yong VW. Promoting oligodendrogenesis and myelin repair using the multiple sclerosis medication glatiramer acetate. Proceedings of the National Academy of Sciences. 2009;106:17992–17997. doi: 10.1073/pnas.0909607106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Sloane JA, Batt C, Ma Y, Harris ZM, Trapp BD, Vartanian T. Hyaluronan blocks oligodendrocyte progenitor maturation and remyelination through TLR2. Proceedings of the National Academy of Sciences. 2010;107:11555–11560. doi: 10.1073/pnas.1006496107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Snow AD, Mar H, Nochlin D, Kimata K, Kato M, Suzuki S, Hassell J, Wight TN. The presence of heparan sulfate proteoglycans in the neuritic plaques and congophilic angiopathy in Alzheimer's disease. Am J Pathol. 1988;133:456–463. [PMC free article] [PubMed] [Google Scholar]
  137. Snow AD, Mar H, Nochlin D, Sekiguchi RT, Kimata K, Koike Y, Wight TN. Early accumulation of heparan sulfate in neurons and in the beta-amyloid protein-containing lesions of Alzheimer”s disease and Down”s syndrome. Am J Pathol. 1990;137:1253–1270. [PMC free article] [PubMed] [Google Scholar]
  138. Sobel RA. The Extracellular Matrix in Multiple Sclerosis Lesions. J Neuropathol Exp Neurol. 1998;57:205. doi: 10.1097/00005072-199803000-00001. [DOI] [PubMed] [Google Scholar]
  139. Sobel RA. The extracellular matrix in multiple sclerosis: an update. Braz. J. Med. Biol. Res. 2001;34:603–609. doi: 10.1590/s0100-879x2001000500007. [DOI] [PubMed] [Google Scholar]
  140. Sobel RA, Ahmed AS. White matter extracellular matrix chondroitin sulfate/dermatan sulfate proteoglycans in multiple sclerosis. J Neuropathol Exp Neurol. 2001;60:1198–1207. doi: 10.1093/jnen/60.12.1198. [DOI] [PubMed] [Google Scholar]
  141. Sobel RA, Chen M, Maeda A, Hinojoza JR. Vitronectin and integrin vitronectin receptor localization in multiple sclerosis lesions. J Neuropathol Exp Neurol. 1995;54:202–213. doi: 10.1097/00005072-199503000-00007. [DOI] [PubMed] [Google Scholar]
  142. Sobel RA, Mitchell ME. Fibronectin in multiple sclerosis lesions. Am J Pathol. 1989;135:161–168. [PMC free article] [PubMed] [Google Scholar]
  143. Soleman S, Yip PK, Duricki DA, Moon LDF. Delayed treatment with chondroitinase ABC promotes sensorimotor recovery and plasticity after stroke in aged rats. Brain. 2012;135:1210–1223. doi: 10.1093/brain/aws027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Starkey ML, Bartus K, Barritt AW, Bradbury EJ. Chondroitinase ABC promotes compensatory sprouting of the intact corticospinal tract and recovery of forelimb function following unilateral pyramidotomy in adult mice. Eur J Neurosci. 2012;36:3665–3678. doi: 10.1111/ejn.12017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Stoffels JMJ, de Jonge JC, Stancic M, Nomden A, van Strien ME, Ma D, Šišková Z, Maier O, ffrench-Constant C, Franklin RJM, Hoekstra D, Zhao C, Baron W. Fibronectin aggregation in multiple sclerosis lesions impairs remyelination. Brain. 2013;136:116–131. doi: 10.1093/brain/aws313. [DOI] [PubMed] [Google Scholar]
  146. Talbott JF, Loy DN, Liu Y, Qiu MS, Bunge MB, Rao MS, Whittemore SR. Endogenous Nkx2.2+/Olig2+ oligodendrocyte precursor cells fail to remyelinate the demyelinated adult rat spinal cord in the absence of astrocytes. Exp Neurol. 2005;192:11–24. doi: 10.1016/j.expneurol.2004.05.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Tan AM, Colletti M, Rorai AT, Skene JHP, Levine JM. Antibodies against the NG2 proteoglycan promote the regeneration of sensory axons within the dorsal columns of the spinal cord. Journal of Neuroscience. 2006;26:4729–4739. doi: 10.1523/JNEUROSCI.3900-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Tan AM, Zhang W, Levine JM. NG2: a component of the glial scar that inhibits axon growth. J Anat. 2005;207:717–725. doi: 10.1111/j.1469-7580.2005.00452.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Tang X, Davies JE, Davies SJA. Changes in distribution, cell associations, and protein expression levels of NG2, neurocan, phosphacan, brevican, versican V2, and tenascin-C during acute to chronic maturation of spinal cord scar tissue. J. Neurosci. Res. 2003;71:427–444. doi: 10.1002/jnr.10523. [DOI] [PubMed] [Google Scholar]
  150. Tester NJ, Howland DR. Chondroitinase ABC improves basic and skilled locomotion in spinal cord injured cats. Exp Neurol. 2008;209:483–496. doi: 10.1016/j.expneurol.2007.07.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Tripathi R, McTigue DM. Prominent oligodendrocyte genesis along the border of spinal contusion lesions. Glia. 2007;55:698–711. doi: 10.1002/glia.20491. [DOI] [PubMed] [Google Scholar]
  152. Turner DC, Flier LA, Carbonetto S. Identification of a cell-surface protein involved in PC12 cell-substratum adhesion and neurite outgrowth on laminin and collagen. J Neurosci. 1989;9:3287–3296. doi: 10.1523/JNEUROSCI.09-09-03287.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Ulrich J, Bornstein MB. Experimental allergic encephalomyelitis (EAE): delayed myelination-inhibition in vitro with EAE-serum: changes in vulnerability of oligodendroglia and newly formed myelin sheaths. Acta Neuropathol. 1973;25:138–148. doi: 10.1007/BF00687558. [DOI] [PubMed] [Google Scholar]
  154. van Horssen J, Bø L, Dijkstra CD, de Vries HE. Extensive extracellular matrix depositions in active multiple sclerosis lesions. Neurobiol Dis. 2006;24:484–491. doi: 10.1016/j.nbd.2006.08.005. [DOI] [PubMed] [Google Scholar]
  155. Verna JM, Fichard A, Saxod R. Influence of glycosaminoglycans on neurite morphology and outgrowth patterns in vitro. Int. J. Dev. Neurosci. 1989;7:389–399. doi: 10.1016/0736-5748(89)90060-9. [DOI] [PubMed] [Google Scholar]
  156. Villanova M, Sewry C, Malandrini A, Toti P, Muntoni F, Merlini L, Torelli S, Tosi P, Maraldi NM, Guazzi GC. Immunolocalization of several laminin chains in the normal human central and peripheral nervous system. J Submicrosc Cytol Pathol. 1997;29:409–413. [PubMed] [Google Scholar]
  157. Wang P-S, Wang J, Xiao Z-C, Pallen CJ. Protein-tyrosine phosphatase alpha acts as an upstream regulator of Fyn signaling to promote oligodendrocyte differentiation and myelination. Journal of Biological Chemistry. 2009;284:33692–33702. doi: 10.1074/jbc.M109.061770. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Wang P-S, Wang J, Zheng Y, Pallen CJ. Loss of protein-tyrosine phosphatase α (PTPα) increases proliferation and delays maturation of oligodendrocyte progenitor cells. Journal of Biological Chemistry. 2012;287:12529–12540. doi: 10.1074/jbc.M111.312769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Watkins TA, Ben Emery, Mulinyawe S, Barres BA. Distinct stages of myelination regulated by gamma-secretase and astrocytes in a rapidly myelinating CNS coculture system. Neuron. 2008;60:555–569. doi: 10.1016/j.neuron.2008.09.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Williams A, Piaton G, Lubetzki C. Astrocytes--friends or foes in multiple sclerosis? Glia. 2007;55:1300–1312. doi: 10.1002/glia.20546. [DOI] [PubMed] [Google Scholar]
  161. Yamaguchi Y. Lecticans: organizers of the brain extracellular matrix. Cell Mol Life Sci. 2000;57:276–289. doi: 10.1007/PL00000690. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Yang Z, Suzuki R, Daniels SB, Brunquell CB, Sala CJ, Nishiyama A. NG2 glial cells provide a favorable substrate for growing axons. Journal of Neuroscience. 2006;26:3829–3839. doi: 10.1523/JNEUROSCI.4247-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Yiu G, He Z. Glial inhibition of CNS axon regeneration. Nat Rev Neurosci. 2006;7:617–627. doi: 10.1038/nrn1956. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Yurchenco PD, Cheng YS, Colognato H. Laminin forms an independent network in basement membranes. J Cell Biol. 1992;117:1119–1133. doi: 10.1083/jcb.117.5.1119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Zhao C, Fancy SPJ, ffrench-Constant C, Franklin RJM. Osteopontin is extensively expressed by macrophages following CNS demyelination but has a redundant role in remyelination. Neurobiol Dis. 2008;31:209–217. doi: 10.1016/j.nbd.2008.04.007. [DOI] [PubMed] [Google Scholar]
  166. Zhao C, Fancy SPJ, Franklin RJM, ffrench-Constant C. Up-regulation of oligodendrocyte precursor cell alphaV integrin and its extracellular ligands during central nervous system remyelination. J. Neurosci. Res. 2009;87:3447–3455. doi: 10.1002/jnr.22231. [DOI] [PubMed] [Google Scholar]
  167. Zhou J, Nagarkatti P, Zhong Y, Nagarkatti M. Immune modulation by chondroitin sulfate and its degraded disaccharide product in the development of an experimental model of multiple sclerosis. J Neuroimmunol. 2010;223:55–64. doi: 10.1016/j.jneuroim.2010.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Zimmermann DR, Dours-Zimmermann MT. Extracellular matrix of the central nervous system: from neglect to challenge. 2008;130:635–653. doi: 10.1007/s00418-008-0485-9. [DOI] [PubMed] [Google Scholar]

RESOURCES