Skip to main content
Journal of Anatomy logoLink to Journal of Anatomy
. 2004 Jan;204(1):33–48. doi: 10.1111/j.1469-7580.2004.00261.x

Chondroitin sulphate proteoglycans: preventing plasticity or protecting the CNS?

KE Rhodes 1, JW Fawcett 1
PMCID: PMC1571240  PMID: 14690476

Abstract

It is well established that axonal regeneration in the adult CNS is largely unsuccessful. Numerous axon-inhibitory molecules are now known to be present in the injured CNS, and various strategies for overcoming these obstacles and enhancing CNS regeneration have been experimentally developed. Recently, the use of chondroitinase-ABC to treat models of CNS injury in vivo has proven to be highly beneficial towards regenerating axons, by degrading the axon-inhibitory chondroitin sulphate glycosaminsoglycan chains found on many proteoglycans in the astroglial scar. This enzyme has now been shown to restore synaptic plasticity in the visual cortex of adult rats by disrupting perineuronal nets, which contain high levels of chondroitin sulphate proteoglycans (CS-PGs) and are expressed postnatally around groups of certain neurons in the normal CNS. The findings suggest exciting prospects for enhancing growth and plasticity in the adult CNS; however, some protective roles of CS-PGs in the CNS have also been demonstrated. Clearly many questions concerning the mechanisms regulating expression of extracellular matrix molecules in CNS pathology remain to be answered.

Keywords: astrogliosis, blood–brain barrier, critical period, spinal cord injury, tenascin

Introduction

Axonal regeneration is poor in the adult CNS

Injured neurons of the adult mammalian central nervous system (CNS) have a limited capacity to regenerate their axons, which rarely succeed in reaching the target and in forming new functional synapses. Consequently, damage to the CNS is normally irreparable. By contrast, axons of the injured peripheral nervous system (PNS) have a strong capacity for regeneration and functional recovery.

What is the reason for this dichotomy? Successful axonal regeneration depends on several factors: survival of the cell body; axonal sprouting from the proximal stump of the lesioned neuron; elongation of these axonal sprouts towards their original target; axonal ensheathment; removal of degenerating axons and eventually restoration of contacts between axon terminals and their target. Whilst these events occur successfully in the PNS (see Bray et al. 1981 for a review), with structural and functional recovery of the nerve, lesioned axons of the CNS generally exhibit only limited sprouting; these sprouts do not elongate and regeneration fails.

A combination of several causes is believed to underlie this failure. CNS axons have an intrinsically reduced capacity for growth: embryonic or early postnatal axons are considerably more able to regenerate following a lesion than those of adults (Kalil & Reh, 1982; Schreyer & Jones, 1983). The CNS environment is clearly less conducive to axonal growth than that of the PNS; the glia of the two environments show a number of differences. The PNS is rich in Schwann cells, and axons from both PNS and CNS neurons grow effectively in such an environment (Richardson et al. 1980, 1984; David & Aguayo, 1981; for review see Aguayo et al. 1981). The CNS, however, contains astrocytes and oligodendrocytes through which both PNS and CNS axons grow poorly (Fawcett et al. 1989a, b; Bandtlow et al. 1990), implying glia and/or other elements of the adult CNS environment are not suitable for axonal regeneration – owing to expression of inhibitory molecules and/or insufficient expression of guidance and trophic factors. Additionally, after damage to the CNS tissue, axonal and glial debris remain for many months and could be a source of axon inhibition. By contrast, debris is cleared from lesions much more rapidly in the PNS (Perry et al. 1987; Stoll et al. 1989).

A number of studies have been carried out over the last two decades to isolate molecular differences between the CNS and PNS environments; numerous potentially axon-inhibitory molecules are found in the CNS (discussed below). Furthermore, where the adult CNS is injured a scar is formed: it is in this scar tissue that axon regeneration fails (Davies et al. 1997). Many of the axon growth-inhibitory molecules have been shown to be up-regulated or only present in the astroglial scar of the lesioned CNS. This astroglial scar is specific to the mature CNS (Berry et al. 1983), and represents a major impediment to growing axons. Although the scar takes days or weeks to mature, it begins to form very rapidly after injury, resulting in a densely packed, highly complex structure consisting of reactive glial cells and up-regulated extracellular matrix (ECM; see below), which surrounds the lesion site.

Growth-inhibitory molecules in the CNS

A variety of molecules known to exist in the adult CNS have been shown through in vitro and in vivo studies to demonstrate axonal growth-inhibitory properties. More importantly, many of these molecules are specific to, or up-regulated in, the injured CNS environment. The molecules may be loosely classified into two main types: myelin-associated molecules and ECM constituents.

Myelin-associated molecules

In the late 1980s Schwab and colleagues demonstrated that a substrate non-permissive to growth exists in CNS white matter (Schwab & Thoenen, 1985; Schwab & Caroni, 1988; Crutcher, 1989; Savio & Schwab, 1989), leading to the hypothesis that where myelinated axons are disrupted, debris containing myelin-associated axon-inhibitory molecules will be present around the lesion. A number of myelin-associated molecules have since been isolated that have axon-inhibitory properties, including the two neurite growth inhibitors Nogo (originally called NI-250) and myelin-associated glycoprotein (MAG). Much literature has been published recently concerning MAG, Nogo and the Nogo receptor (Nogo-66) that is beyond the scope of this article; for recent reviews see Bandtlow (2003) and McGee & Strittmatter (2003).

ECM constituents

For many neurons, their migration and axon elongation occurs through the ECM: in the CNS the ECM is largely devoid of cells but contains several types of molecules with which neurons and glia interact, and these can have important influences on many aspects of a cell's behaviour.

The extracellular space surrounding many non-neuronal cells in the CNS is filled with a network of glycoproteins, proteoglycans and hyaluronan; close to the membrane of such cells the ECM becomes more dense, forming a basement membrane composed principally of collagens; glycoproteins – particularly tenascin-C and tenascin-R; chondroitin sulphate proteoglycans (CS-PGs) and heparan sulphate proteoglycans (HS-PGs); hyaluronic acid (HA); cell adhesion molecules and integrins.

ECM molecules expressed in the developing CNS may have both growth-promoting and growth-inhibitory effects on axons (e.g. Snow et al. 1990b; Oakley & Tosney, 1991; Brittis et al. 1992; Emerling & Lander, 1996; Treloar et al. 1996; for review see Margolis & Margolis, 1997). Many of these molecules are up-regulated in the adult ECM following a lesion to the CNS (in particular CS-PGs; see below), and have been shown to have inhibitory properties towards regenerating axons.

Tenascin is an important secreted ECM component with a range of binding sites and functions (Hoffman et al. 1988; Steindler et al. 1989; Grumet et al. 1994; Husmann et al. 1995; for review see Faissner, 1997). Tenascin is abundant in the basement membrane, being produced by astrocytes during development, with important roles in mediating axon–glia interactions (Steindler et al. 1989; Faissner & Kruse, 1990; Lochter et al. 1991). There are two members of the tenascin gene family: tenascin-C and tenascin-R; tenascin-C is expressed as numerous alternatively spliced variants with various functions (Faissner et al. 1988; Stern et al. 1989; Chuong & Chen, 1991; Faissner, 1997). The same tenascin molecule may have either growth-inhibitory or growth-promoting effects towards different neurons within different contexts; a number of studies have demonstrated the neurite growth-inhibitory properties of tenascin in vitro (Pesheva et al. 1989; Crossin et al. 1990; Faissner & Kruse, 1990). It also has growth-promoting effects ascribed to the alternatively spliced A-D and D5 domains (Meiners et al. 1999, 2001).

Tenascin is found in the normal adult CNS although at lower levels than in the developing CNS. Production is up-regulated in the glial scar after injury (Laywell et al. 1992); the increased tenascin has been co-localized with reactive glial fibrillary acidic protein (GFAP)+ astrocytes (Lochter et al. 1991; McKeon et al. 1991; Laywell et al. 1992). Tenascin is known to interact with many CS-PGs in vitro (Grumet et al. 1994; Xiao et al. 1997, 1998), and thus may be capable of forming (axon-inhibitory) complexes with CS-PGs in injured CNS tissue.

The astroglial scar

A lesion to the CNS provokes reactive behaviour among local glia, which proliferate and/or migrate to the lesion site, forming the astroglial scar (Fig. 1). These cells – principally astrocytes, microglia and oligodendrocyte precursor (OP) cells – respond rapidly: they may be identified by their hypertrophied cell body and thickened processes, and altered production of a number of cell surface receptors and ECM molecules, including tenascin and proteoglycans (McKeon et al. 1991, 1995; Levine, 1994; Haas et al. 1999; Asher et al. 2000; see reviews by Streit & Kincaid-Colton, 1995; Ridet et al. 1997; Raivich et al. 1999). Reactive microglia and OP cells also undergo local proliferation: these cells are recruited around the lesion site in large numbers (Levine, 1994; Rhodes et al. 2003), where they contribute to the astroglial scar. These changes begin within hours and develop over about 7–10 days, leading eventually to the classic appearance of the mature glial scar, which consists of highly branched astrocytic processes attached to one another by junctional processes, with plentiful ECM (Maxwell et al. 1990), surrounding macrophages, fibroblasts and reactive glia at the lesion core.

Fig. 1.

Fig. 1

Simplified diagram showing major functions, markers expressed and molecules produced by glial cells in (A) the normal adult CNS and (B) the lesioned CNS (direct trauma). Morphological, phenotypical and functional changes are observed among glial cells contributing to glial scar formation around the lesion site; oligodendrocytes are not included. Inflammatory cells (monocytes, neutrophils and T cells) would also be recruited across the compromised blood–brain barrier (not illustrated), producing a range of pro-inflammatory chemotactic cytokines and chemokines, assisting endogenous microglia with antigen recognition and phagocytotic functions, and stimulating further gliosis. Bold type indicates new expression of a protein; ↑ = up-regulated expression of a protein; scale bars = 5 µm. Images taken from adult rat basal ganglia in normal brain (A) and 24 h following microinjection of 5 µg lipopolysaccharide (B) causing rapid inflammatory response and gliosis in microglia (labelled with antibodies to CR3), astrocytes (labelled with antibodies to GFAP) and OP cells (labelled with antibodies to NG2). Abbreviations not in text: CAM, cell adhesion molecule; CR3, complement receptor 3; FGF, fibroblast growth factor; MCSF(-R), macrophage colony stimulating factor (receptor); MHC, major histocompatibility complex; NGF, nerve growth factor; PDGFαR, platelet-derived growth factor receptor-alpha; TGF, transforming growth factor; TNF, tumour necrosis factor. Based on information from Ridet et al. (1997), Raivich et al. (1999) and Asher et al. (2001).

Chondroitin sulphate proteoglycans

Significant studies by Davies et al. (1997, 1999) demonstrated that adult dorsal root ganglion (DRG) neurons microtransplanted into undamaged adult white matter tracts were able to grow through this CNS environment. However, these axons stopped growing upon reaching damaged CNS tissue. Immunohistochemistry revealed that the ECM-rich gliotic scar surrounding the damaged tissue contained a large proportion of CS-PGs, the distribution of these molecules matching the regions of cessation of axon growth. These and other studies suggest that glial scar tissue is more inhibitory to axon regeneration than undamaged adult CNS tissue, and show a correlation between inhibition of axon growth and the presence of elevated levels of CS-PGs.

Proteoglycans: general structure

Proteoglycans encompass a wide range of complex molecules found throughout different tissue types; they have been associated with a variety of cellular processes including cell adhesion, growth, receptor binding, cell migration, barrier formation and interaction with other ECM constituents (for review see Bandtlow & Zimmerman, 2000). These diverse molecules comprise a core glycoprotein with glycosaminoglycan (GAG) sugar chains covalently attached. Each GAG consists of a simple, linear polymer of repeating disaccharide units, composed from two alternating monosaccharides: usually one sugar is a uronic acid and the other is either N-acetylglucosamine or N-acetylgalactosamine. Different types of GAG may be created as a result of sulphation and epimerization modifications that are carried out on the sugars themselves following polymerization. The length of GAG chains may also vary, from a polypeptide chain of just 10 kDa to 400 000 kDa, and the core protein itself may have varying numbers of GAG chains attached – from one to well over 100, resulting in a diverse array of PGs with a range of functional complexities. Many of the functional properties of PGs are attributed to the attached side chains; much of the interaction between PGs and cell-surface receptors or ECM proteins is thought to occur via binding sites on the GAG chains, although the core protein is also able to bind substrates (see Bandtlow & Zimmerman, 2000).

CS-PG expression patterns and functions in the CNS

The CNS is rich in proteoglycans, particularly in those containing sulphated GAG, such as CS-, HS-, dermatan sulphate (DS)- and keratan sulphate (KS)-PGs. CS-PGs are characterized by sulphation of the GAG side chains at certain positions on the sugars; these GAG chains may be cleaved from the core protein by chondroitinase-ABC (or chondroitinase-AC).

The CS-PGs are represented by several types, including large aggregating proteoglycans such as aggrecan (Paulsson et al. 1987) and versican (Krusius et al. 1987); the brain-specific proteoglycans neurocan (Grumet et al. 1993, 1994; Friedlander et al. 1994; Oohira et al. 1994) and brevican (Yamada et al. 1994); and NG2 (Levine & Card, 1987; Stallcup & Beasley, 1987) and phosphacan/DSD-1 (Grumet et al. 1993, 1994; Maeda et al. 1994; Maurel et al. 1994). These CS-PGs are abundant within the ECM; all are expressed in the CNS, and have been shown to interact with a variety of other matrix components including laminin, fibronectin (Schmidt et al. 1991), tenascin (Grumet et al. 1994), HA (Krusius et al. 1987; Paulsson et al. 1987; Doege et al. 1991) and collagen (Bidanset et al. 1992; Hedbom & Heinegard, 1993). Interactions of CS-PGs with cell surface receptors or other ECM components may occur through the GAG chains or via the core protein.

CS-PG expression is particularly strong in the embryonic brain: the patterns of expression during development (in regions such as neural crest pathways and spinal cord) (Snow et al. 1990a; Oakley & Tosney, 1991) and axon-inhibitory properties are consistent with the theory that CS-PGs behave as axon guidance molecules during development, because growing axons generally avoid regions rich in CS-PGs (Snow et al. 1990b; Oakley & Tosney, 1991; for review see Asher et al. 2001), although there are some reports of axon-promoting mechanisms of CS-PGs (Brittis et al. 1992; Emerling & Lander, 1996; for review see Margolis & Margolis, 1997). Furthermore, it has been demonstrated in vitro that removal of chondroitin sulphate from models of developmental axon guidance abolishes the inhibition (e.g. Brittis et al. 1992).

The expression patterns of CS-PGs are altered in the adult CNS. For example, neurocan is mainly present in the white matter in the adult CNS; this protein is produced largely by astrocytes but can also be produced by OP cells (Asher et al. 2000). NG2 is distributed throughout the adult brain on OP cells (Levine & Card, 1987; Levine et al. 1993), on some blood vessels and meningeal cells (Morgenstern et al. 2002); this molecule may additionally be cleaved and secreted into the ECM (Nishiyama et al. 1995). Phosphacan/DSD-1 and the membrane-associated form, RPTPβ, is present at particularly high levels in Bergmann glia and nuclei in the cerebellum of the postnatal rat, and elsewhere in the brain and spinal cord (Meyer-Puttlitz et al. 1996). Versican is found in the white matter, and produced by OP cells (Asher et al. 2002); brevican may be found distributed throughout the CNS (Seidenbecher et al. 1995), produced by astrocytes (Yamada et al. 1997). Versican, brevican and aggrecan are all thought to be expressed in greater quantities in the normal adult brain than in the developing brain (Milev et al. 1998).

CS-PGs inhibit neurite growth in vitro and in vivo

Many studies have demonstrated in vitro that CS-PGs are inhibitory towards neurite outgrowth, either via their CS chains or their core proteins (Dou & Levine, 1994, 1997; Friedlander et al. 1994; Smith-Thomas et al. 1994; Yamada et al. 1997; Niederöst et al. 1999; Asher et al. 2000; Schmalfeldt et al. 2000). Additionally, both phosphacan and neurocan are able to bind with high affinity to the cell adhesion molecules N-CAM and Ng-CAM/L1 (Grumet et al. 1993; Friedlander et al. 1994), thus interfering with their interactions and indirectly inhibiting neurite outgrowth (Friedlander et al. 1994; Milev et al. 1994).

Reactive astrocytes and/or OP cells within the glial scar (see Fig. 1) have been shown to up-regulate their expression of various axon-inhibitory CS-PGs in vivo, in particular neurocan (Haas et al. 1999; McKeon et al. 1999; Asher et al. 2000), NG2 (Levine, 1994; Ong & Levine, 1999; Rhodes et al. 2003), phosphacan (Laywell & Steindler, 1991; McKeon et al. 1995) and versican (Asher et al. 2002; Jones et al. 2003; Tang et al. 2003). Several studies have identified an up-regulation of inhibitory CS in CNS lesion sites, using the antibody CS56, which binds to CS chains on proteoglycans (e.g. McKeon et al. 1991; Davies et al. 1997, 1999; Fitch & Silver, 1997).

Overcoming CS-PG inhibition of axon growth: chondroitinase-ABC

It has been established in vitro that much of the axon growth-inhibitory properties of several CS-PGs reside in the CS side chains because their inhibitory effects may be abolished with the enzyme chondroitinase-ABC, or by interfering with CS GAG synthesis (Friedlander et al. 1994; Milev et al. 1994; Smith-Thomas et al. 1995; Schmalfeldt et al. 2000). NG2 is an unusual CS-PG in that the inhibitory effects of the purified molecule on neurite outgrowth are not affected by chondroitinase (Dou & Levine, 1994); however, the use of NG2 neutralizing antibodies has been shown to abolish inhibitory effects of NG2-containing cell membranes (Chen ZJ et al. 2002).

Using an ex vivo model of the glial scar (Rudge et al. 1989), adult scar tissue was shown to be rich in reactive astrocytes and CS-PGs, which were inhibitory towards neurite outgrowth in vitro; treatment with chondroitinase was able to overcome this inhibition (McKeon et al. 1991, 1995).

More recently chondroitinase has been applied to models of CNS injury in vivo. Treatment of the lesioned adult rat nigrostriatal system (Moon et al. 2001) or dorsal column (Bradbury et al. 2002) with chondroitinase degraded CS-GAG around the lesion site and promoted axonal regeneration. Bradbury et al. (2002) demonstrated an accompanying increase in the expression of the growth-associated protein (GAP)-43 in lesioned neurons, restoration of post-synaptic activity below the lesion and functional recovery. The beneficial effects of chondroitinase towards aiding axonal regeneration after injury have been shown in further recent studies (Krekoski et al. 2001; Zuo et al. 2002; Tropea et al. 2003; Yick et al. 2003).

Protective roles of CS-PGs in the CNS

CS-PGs are up-regulated where the blood–brain barrier breaks down

It is currently not known precisely what triggers gliosis and eventual formation of the astroglial scar, the end product of a cascade of events. It has been observed, however, that up-regulation of CS-PGs may be immunohistochemically localized to the region of tissue damage and inflammation, and hence compromise of the blood–brain barrier (BBB) (Fitch & Silver, 1997; Fitch et al. 1999; Rhodes et al. 2002; our unpublished personal observations).

Some studies have suggested candidate molecules involved in promoting gliosis as a result of BBB breakdown: in particular, the serine protease thrombin has been implicated. Thrombin, derived from its zymogen prothrombin and an important part of the blood coagulation cascade (see Davie et al. 1991, for review), has multiple functions including activation of platelets, and chemotaxis and adhesion in monocytes, macrophages and neutrophils (reviewed by Grand et al. 1996). Thrombin may also influence cells in the CNS, including regulating neurite outgrowth (Gurwitz & Cunningham, 1988), inducing astrocytic proliferation (Grabham & Cunningham, 1995) and microglial activation (Möller et al. 2000). Thrombin and other serine proteases would extravasate across a compromised BBB; however, the rapid effects of thrombin on CNS cells demonstrated in vitro are not mimicked by other serine proteases (see Grand et al. 1996). Nishino et al. (1993) infused thrombin into rat brains for 7 days and provoked significant inflammation and gliotic reaction, including infiltration of inflammatory cells; proliferation; reactivity of astrocytes and induction of angiogenesis. In support of these findings, mRNA to both prothrombin and one of the thrombin receptors (belonging to the family of protease activated receptors (PAR) 1–4), PAR-1, is found widely distributed throughout the brain in neurons and glia (Dihanich et al. 1991; Weinstein et al. 1995; Niclou et al. 1998; see review by Gingrich & Traynelis, 2000). Other candidate molecules suggested to trigger gliosis include pro-inflammatory cytokines such as interleukin (IL)-1 (Giulian & Lachman, 1985; Giulian et al. 1988) and ciliary-derived neurotrophic factor (CNTF) (Kahn et al. 1995; Winter et al. 1995). Although all of these molecules clearly play important roles in orchestrating the CNS injury response, none has yet been established as the primary trigger of gliosis; it is likely that a combination of factors is required in order to initiate up-regulation of CS-PGs at sites of tissue damage.

Whilst astrocytes and microglia may become reactive following lesions affecting the CNS where the BBB remains intact (for example around centrally located cell bodies of peripherally lesioned axons, or at the distal or proximal ends of injured neurons; see Raivich et al. 1999), it would appear that such reactive astrocytes do not alter their expression patterns of CS-PGs, although their phenotypes appear otherwise reactive (see Ridet et al. 1997). In keeping with this, Fitch & Silver (1997) demonstrated that increased CS-PG immunoreactivity correlated with evidence of BBB breakdown but not necessarily with astrogliosis further from the lesion. Additionally, observations from within our laboratory suggest that OP cells only become reactive where the BBB is compromised, these cells rapidly increase their expression of NG2, and take on a reactive morphology (Rhodes et al. 2002).

Sobel & Ahmed (2001) correlated increased expression of the CS-PGs aggrecan, neurocan and versican with increased astrocytosis in the pro-inflammatory environment surrounding multiple sclerosis (MS) plaques in human patients; these alterations to the ECM were believed to be a result of the compromised BBB, and ensuing release and activation of proteases. Interestingly, when Fitch and colleagues examined injured rat CNS tissue, it was found that cavities which had developed near to the original wound, as made evident by strong inflammatory reactions, were equally demarcated by up-regulated CS-PG expression (Fitch & Silver, 1997; Fitch et al. 1999).

These findings suggest proteoglycans surrounding regions of damaged tissue and disrupted BBB have a vital role in sealing off the lesioned area and perhaps limiting further cavitation and secondary injury. Such a mechanism would rely on the very rapid ability for local glial cells to increase their production of CS-PGs in response to factors either present in extravasing serum, or on infiltrating inflammatory cells.

Although further understanding of the triggers of increased CS-PG synthesis at sites of CNS tissue damage might enable development of tools that would allow greater manipulation of the axon-inhibitory glial scar, the findings discussed above suggest that preventing up-regulation of CS-PGs at the site of a wound might have detrimental effects on the healing and sealing processes.

Proteoglycans and models of disease

Given the abundancy of proteoglycans in ECM and connective tissues, it is perhaps unsurprising that these molecules have been implicated in a diverse range of syndromes or diseases. Investigations of pathologies of connective tissues (including cartilage and skeletal systems, particularly forms of arthritis) have associated the pathology with degradation of the ECM and a resulting lack of DS-/CS-PGs that are an important component of the articular cartilage (e.g. Richardson, 1985; Pulkkinen et al. 1990; Blackshear et al. 1997; for review see Aigner & McKenna, 2002). More recently, models of experimentally induced proteoglycan deficiencies have linked mutant mice lacking various proteoglycans with abnormal development and pathological phenotypes (e.g. Xu et al. 1998; Serpinskaya et al. 1999; Baeg et al. 2001; Ameye et al. 2002; Chen XD et al. 2002; Jepsen et al. 2002) (for recent reviews see Ameye & Young, 2002; Schwartz & Domowicz, 2002).

There has also been some evidence of neuroprotective roles of CS-PGs within the CNS. For example, recent studies examining the mechanisms of beta amyloid-induced neurodegeneration in models of Alzheimer's disease have noted a correlation between neurons that are normally devoid of perineuronal nets (PNNs; see below) and their susceptibility to Alzheimer's disease (Brückner et al. 1999; Adams et al. 2001; Hartig et al. 2001), suggesting that the high concentration of CS-PGs surrounding neurons with PNNs can somehow protect them from the formation of Alzheimer's lesions. The initiation of neurodegenerative processes has been associated with abnormal deposition of HS-, DS- and CS-PGs and increased proteolytic cleavage of ECM components (Belichenko et al. 1999). Studies in vitro have additionally shown neuroprotective effects of chondroitin sulphate: CS-PGs were added to cultures of rat hippocampal or cortical neurons in order to rescue the cells from excitotoxic damage (Okamoto et al. 1994a, b); Woods et al. (1995) attenuated beta-amyloid-induced neurodegeneration of hippocampal cultures using CS- and HS-GAG. Chondroitin sulphates are also able to protect chondrocytes from inflammation and articular symptoms in models of osteoarthritis both in vitro (Nerucci et al. 2000) and in vivo (Uebelhart et al. 1998); this has been reviewed by Bali et al. (2001).

Pathological implications for HS-PGs

Interestingly, the HS-PGs seem to have been more widely implicated with human pathology than CS-PGs. Whilst the proteoglycan form of amyloid precursor protein (APP), appican, does contain a CS chain (Shioi et al. 1992, 1993) and CS-PGs have been implicated in Alzheimer's disease (e.g. De Witt et al. 1994, 1994), the majority of papers demonstrate increased expression of HS-PGs in neurodegenerative plaques (e.g. Schubert et al. 1988; Snow et al. 1988; Su et al. 1992; Caceres & Brandan, 1997; for review see Bornemann & Staufenbiel, 2000). HS-PGs have also been shown to promote phosphorylation of the microtubule associated protein tau (Goedert et al. 1996); Canning et al. (1993) showed beta amyloid in the brain stimulates astrocytes to become reactive and increase their deposition of CS- or HS-PGs.

Much literature has also associated up-regulation of HS-PGs (syndecans and glypicans in particular) with invasive gliomas (e.g. Steck et al. 1989; Liang et al. 1997; Barbareschi et al. 2003); the CS-PGs brevican (Zhang et al. 1998; Gary et al. 2000; Nutt et al. 2001), versican (Paulus et al. 1996; Ang et al. 1999; Touab et al. 2002) and NG2 (Schrappe et al. 1991; Burg et al. 1997; Chekenya et al. 1999; Shoshan et al. 1999) have also been been implicated.

It would seem that the local balance of PGs in the ECM or pericellular environment is crucial: mechanisms that cause a deficiency (as in arthritic diseases) or abnormal accumulation of HS- and CS-PGs (as in astrocytosis and neurodegenerative lesions) have been associated with damaging pathology. The presence of proteoglycans under normal conditions is clearly necessary and crucial for cellular maintenance: those mechanisms that upset the balance of proteoglycans and their sulphation patterns still need to be elucidated.

Perineuronal nets

Structure and composition

PNNs were first described by Golgi and Cajal in the 1890s as reticular networks observed on the surface of neuronal cell bodies and proximal dendrites. They develop postnatally, finally appearing as net-like features on the cell surface as a result of ECM materials deposited around synaptic endings, and in the space between neurons and astrocytic processes. PNNs were originally visualized using iron- or silver-based histochemical stains; most PNNs can be detected using lectins such as the Wisteria floribunda agglutinin lectin, which recognizes N-acetylgalactosamine in CS-GAG chains (Hartig et al. 1992). Although it is not fully understood exactly which cell types produce the PNN, or why, studies so far have found PNNs most commonly around parvalbumin-containing GABAergic interneurons (e.g. Hartig et al. 1992, 1994; Brückner et al. 1994; Murakami et al. 1995) and pyramidal cells (Hausen et al. 1996) in the cortex, and around projection neurons and large motorneurons of the brain stem and spinal cord (70–80% neurons in the cord have PNNs) (Asher et al. 1995; Murakami et al. 1995; Takahashi-Iwanaga et al. 1998). Because glial cells in vitro can produce ECM similar to PNNs observed in vivo, even in the absence of neurons (Maleski & Hockfield, 1997), it seems likely that astrocytes contribute at least in part to PNN formation. Interestingly, in the culture conditions utilized by Maleski & Hockfield (1997), neurons were never seen to assemble any kind of pericellular matrix – although there is a possibility that this was an artefact resulting from the conditions because Knudson & Knudson (1991) demonstrated that some cells in vitro never produce pericellular matrices, despite having HA receptors and the ability to assemble PNN-like matrix from exogenously added HA and aggrecan.

The structure and composition of PNNs has become more widely understood in recent years. Asher et al. (1995) demonstrated immunohistochemically the existence of major cartilage constituents in the perineuronal matrix of bovine spinal cord: aggrecan and link protein, both of which were hyaluronate-dependent as their immunoreactivities could be abolished by pretreating with the enzyme hyaluronidase. Aggrecan is a large proteoglycan belonging to the lectican family of CS-PGs, which also includes brevican, neurocan and versican. Whereas aggrecan and versican are present in a number of connective tissues (e.g. Mundlos et al. 1991; Yamagata et al. 1993; Glumoff et al. 1994; Popp et al. 2003), brevican and neurocan are usually found only in neural tissues (Rauch et al. 1992; Oohira et al. 1994; Yamada et al. 1994; Seidenbecher et al. 1995; Watanabe et al. 1995). All four of these lecticans have since been detected in PNNs (Bignami et al. 1993; Koppe et al. 1997a, b; Brückner et al. 1998; Matsui et al. 1998; Matthews et al. 2002), as has phosphacan/DSD-1 (Wintergerst et al. 1996; Haunso et al. 1999) and tenascin-R (Brückner et al. 2000; Haunso et al. 2000). In a recent review, Yamaguchi (2000) proposed a model of PNNs whereby long HA molecules bind CS-PGs (lecticans), which in turn bind tenascin and form net-like complexes between neurons and glia.

PNNs are not all identical in composition: whereas PNNs in the rat deep cerebellar nuclei and spinal cord, for example, are rich in expression of phosphacan, simultaneously labelled sections of cortex demonstrate little phosphacan (Fig. 2). Antibodies to the aggrecan-like antigen CAT-301 appear to label PNNs as frequently as the Wisteria floribunda lectin, whereas versican expression is more distinguishable around large motorneurons in the spinal cord and brainstem. The importance of differential expression patterns of CS-PGs in PNNs is unclear; however, tenascin appears to be a more critical component, because mutant tenascin-knockout mice showed disrupted PNNs (Weber et al. 1999; Brückner et al. 2000; Haunso et al. 2000), whereas mice deficient in neurocan (Zhou et al. 2001) or brevican (Brakebusch et al. 2002) were anatomically and morphologically similar to wild-type mice.

Fig. 2.

Fig. 2

Immunohistochemistry demonstrating PNNs in 4% paraformadlehyde-perfused adult rat CNS using Wisteria floribunda agglutinin lectin (WFAL; A–C); antibodies to phosphacan (2B49; D–F); and antibodies to versican (12C5; G–I). PNNs are clearly demarcated with WFAL around small neurons in the cortex (A) and larger motorneurons in deep cerebellar nuclei (DCN; B) and spinal cord (C). Phosphacan expression in cortex (D) is faint with few distinguishable PNNs, although strong perineuronal labelling is seen in DCN (E) and spinal cord (F). Strong expression of versican throughout the cortex (G) makes it harder to distinguish individual PNNs, although perineuronal expression of versican appears stronger in DCN (H) and spinal cord (I). Scale bars = 50 µm except (C) 10 µm; arrows point to examples of perineuronal nets.

Functional roles for CS-PGs and PNNs in synaptic plasticity

The late development of PNNs, and their association with only certain neuronal types, has led to speculation of their functions including roles in maintenance of tissue architecture (Margolis & Margolis, 1993; Celio & Blumcke, 1994) and involvement in maturation of synapses on motorneurons (Kalb & Hockfield, 1988, 1990). Following the observation that neuronal surfaces covered by PNNs are devoid of synaptic contacts (Brückner et al. 1993), it has been proposed that PNNs may specifically serve to block formation of new synapses (Celio & Blumcke, 1994; for review see Yamaguchi 2000).

It has been shown that the late postnatal appearance of PNNs in the visual cortex coincides with the end of the so-called critical period (Sur et al. 1988; Zaremba et al. 1989; Koppe et al. 1997a, b), a well-documented developmental time-window during which synapses are still open to experience (e.g. Pettigrew, 1972; Daw & Wyatt, 1976) (for a review of synaptic plasticity and ECM molecules see Dityatev & Schachner, 2003). At the end of the critical period, synaptic plasticity is greatly diminished – however, visually deprived (dark-reared) rats retain plasticity of the visual cortex (Cynader et al. 1976): the phenomenon has been well characterized (recently reviewed by Berardi et al. 2003). Studies from Hockfield and colleagues previously demonstrated changes in expression of CS-PGs in the visual cortex in visually deprived rats (Zaremba et al. 1989; Guimaraes et al. 1990; Hockfield et al. 1990), implying a role for CS-PGs and PNNs in maintaining synaptic stability and preventing plasticity in the mature animal (for review see Hockfield & Kalb, 1993).

In support of these earlier studies, Pizzorusso et al. (2002) demonstrated that treatment of the mature rat visual cortex with chondroitinase ABC degrades CS-PGs present in the PNNs, restoring synaptic plasticity normally only seen during the critical period in the development, or in dark-reared animals. Furthermore, mutant mice lacking neurocan (Zhou et al. 2001) or brevican (Brakebusch et al. 2002) were shown to have defects in hippocampal long-term potentiation although their brains were anatomically and morphologically indistinguishable from those of wild-type mice. Bukalo et al. (2001) found differential effects of removal of tenascin-R (using knockout mice) or CS-PGs (using chondroitinase-ABC) on hippocampal synaptic plasticity. These findings have important implications for the future, showing that it may be possible to restore growth potential and plasticity to the mature CNS.

Conclusions

The mature CNS is highly prohibitive to establishment of new synaptic connections or regeneration of injured neurons. Research over the past few decades has highlighted many molecules (in particular, CS-PGs) that appear to be involved in both maturation of synapses and prevention of new axonal growth. Although it must be assumed that these mechanisms exist in order to prevent aberrant neuronal growth and accidental rewiring of the CNS, recent advances have demonstrated that manipulation of the environment using substances such as chondroitinase-ABC can override these barriers. The findings suggest exciting prospects for the future, particularly for therapeutic research in overcoming debilitating spinal cord injury, and increasing CNS plasticity; however, many questions remain to be answered. Some protective roles for CS-PGs have been highlighted; it is clear that abnormal deficiency or accumulation of CS-PGs leads to pathological states; further investigation of the long-term effects of manipulating the balance of ECM molecules in the CNS needs to be conducted.

Acknowledgments

The authors are funded by the International Spinal Research Trust, the Medical Research Council and the Wellcome Trust.

References

  1. Adams I, Brauer K, Arelin C, Hartig W, Fine A, Mader M, et al. Perineuronal nets in the rhesus monkey and human basal forebrain including basal ganglia. Neuroscience. 2001;108:285–298. doi: 10.1016/s0306-4522(01)00419-5. [DOI] [PubMed] [Google Scholar]
  2. Aguayo AJ, David S, Bray GM. Influences of the glial environment on the elongation of axons after injury: transplantation studies in adult rodents. J. Exp. Biol. 1981;95:231–240. doi: 10.1242/jeb.95.1.231. [DOI] [PubMed] [Google Scholar]
  3. Aigner T, McKenna L. Molecular pathology and pathobiology of osteoarthritic cartilage. Cell Mol. Life Sci. 2002;59:5–18. doi: 10.1007/s00018-002-8400-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Ameye L, Aria D, Jepsen K, Oldberg A, Xu T, Young MF. Abnormal collagen fibrils in tendons of biglycan/fibromodulin-deficient mice lead to gait impairment, ectopic ossification, and osteoarthritis. FASEB J. 2002;16:673–680. doi: 10.1096/fj.01-0848com. [DOI] [PubMed] [Google Scholar]
  5. Ameye L, Young MF. Mice deficient in small leucine-rich proteoglycans: novel in vivo models for osteoporosis, osteoarthritis, Ehlers–Danlos syndrome, muscular dystrophy, and corneal diseases. Glycobiology. 2002;12:107R–116R. doi: 10.1093/glycob/cwf065. [DOI] [PubMed] [Google Scholar]
  6. Ang LC, Zhang Y, Cao L, Yang BL, Young B, Kiani C, et al. Versican enhances locomotion of astrocytoma cells and reduces cell adhesion through its G1 domain. J. Neuropathol. Exp. Neurol. 1999;58:597–605. doi: 10.1097/00005072-199906000-00004. [DOI] [PubMed] [Google Scholar]
  7. Asher RA, Scheibe RJ, Keiser HD, Bignami A. On the existence of a cartilage-like proteoglycan and link proteins in the central nervous system. Glia. 1995;13:294–308. doi: 10.1002/glia.440130406. [DOI] [PubMed] [Google Scholar]
  8. Asher RA, Morgenstern DA, Fidler PS, Adcock KH, Oohira A, Braistead JE, et al. Neurocan is upregulated in injured brain and in cytokine-treated astrocytes. J. Neurosci. 2000;20:2427–2438. doi: 10.1523/JNEUROSCI.20-07-02427.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Asher RA, Morgenstern DA, Moon LD, Fawcett JW. Chondroitin sulphate proteoglycans: inhibitory components of the glial scar. Prog. Brain Res. 2001;132:611–619. doi: 10.1016/S0079-6123(01)32106-4. [DOI] [PubMed] [Google Scholar]
  10. Asher RA, Morgenstern DA, Shearer MC, Adcock KH, Pesheva P, Fawcett JW. Versican is upregulated in CNS injury and is a product of oligodendrocyte lineage cells. J. Neurosci. 2002;22:2225–2236. doi: 10.1523/JNEUROSCI.22-06-02225.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Baeg GH, Lin X, Khare N, Baumgartner S, Perrimon N. Heparan sulfate proteoglycans are critical for the organization of the extracellular distribution of Wingless. Development. 2001;128:87–94. doi: 10.1242/dev.128.1.87. [DOI] [PubMed] [Google Scholar]
  12. Bali JP, Cousse H, Neuzil E. Biochemical basis of the pharmacologic action of chondroitin sulfates on the osteoarticular system. Semin. Arthritis Rheum. 2001;31:58–68. doi: 10.1053/sarh.2000.24874. [DOI] [PubMed] [Google Scholar]
  13. Bandtlow C, Zachleder T, Schwab ME. Oligodendrocytes arrest neurite growth by contact inhibition. J. Neurosci. 1990;10:3837–3848. doi: 10.1523/JNEUROSCI.10-12-03837.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Bandtlow CE, Zimmermann DR. Proteoglycans in the developing brain: new conceptual insights for old proteins. Physiol. Rev. 2000;80:1267–1290. [Google Scholar]
  15. Bandtlow CE. Regeneration in the central nervous system. Exp. Gerontol. 2003;38:79–86. doi: 10.1016/s0531-5565(02)00165-1. [DOI] [PubMed] [Google Scholar]
  16. Barbareschi M, Maisonneuve P, Aldovini D, Cangi MG, Pecciarini L, Angelo MF, et al. High syndecan-1 expression in breast carcinoma is related to an aggressive phenotype and to poorer prognosis. Cancer. 2003;98:474–483. doi: 10.1002/cncr.11515. [DOI] [PubMed] [Google Scholar]
  17. Belichenko PV, Miklossy J, Belser B, Budka H, Celio MR. Early destruction of the extracellular matrix around parvalbumin-immunoreactive interneurons in Creutzfeldt-Jakob disease. Neurobiol. Dis. 1999;6:269–279. doi: 10.1006/nbdi.1999.0245. [DOI] [PubMed] [Google Scholar]
  18. Berardi N, Pizzorusso T, Ratto GM, Maffei L. Molecular basis of plasticity in the visual cortex. Trends Neurosci. 2003;26:369–378. doi: 10.1016/S0166-2236(03)00168-1. [DOI] [PubMed] [Google Scholar]
  19. Berry M, Maxwell WL, Logan A, Mathewson A, McConnell P, Ashhurst DE, et al. Deposition of scar tissue in the central nervous system. Acta Neurochir. (Suppl.) (Wien.) 1983;32:31–53. doi: 10.1007/978-3-7091-4147-2_3. [DOI] [PubMed] [Google Scholar]
  20. Bidanset DJ, Guidry C, Rosenberg LC, Choi HU, Timpl R, Hook M. Binding of the proteoglycan decorin to collagen type VI. J. Biol. Chem. 1992;267:5250–5256. [PubMed] [Google Scholar]
  21. Bignami A, Perides G, Rahemtulla F. Versican, a hyaluronate-binding proteoglycan of embryonal precartilaginous mesenchyma, is mainly expressed postnatally in rat brain. J. Neurosci. Res. 1993;34:97–106. doi: 10.1002/jnr.490340110. [DOI] [PubMed] [Google Scholar]
  22. Blackshear PJ, Silver J, Nairn AC, Sulik KK, Squier MV, Stumpo DJ, et al. Widespread neuronal ectopia associated with secondary defects in cerebrocortical chondroitin sulfate proteoglycans and basal lamina in MARCKS-deficient mice. Exp. Neurol. 1997;145:46–61. doi: 10.1006/exnr.1997.6475. [DOI] [PubMed] [Google Scholar]
  23. Bornemann KD, Staufenbiel M. Transgenic mouse models of Alzheimer's disease. Ann. NY Acad. Sci. 2000;908:260–266. doi: 10.1111/j.1749-6632.2000.tb06653.x. [DOI] [PubMed] [Google Scholar]
  24. Bradbury EJ, Moon LD, Popat RJ, King VR, Bennett GS, Patel PN, et al. Chondroitinase ABC promotes functional recovery after spinal cord injury. Nature. 2002;416:636–640. doi: 10.1038/416636a. [DOI] [PubMed] [Google Scholar]
  25. Brakebusch C, Seidenbecher CI, Asztely F, Rauch U, Matthies H, Meyer H, et al. Brevican-deficient mice display impaired hippocampal CA1 long-term potentiation but show no obvious deficits in learning and memory. Mol. Cell Biol. 2002;22:7417–7427. doi: 10.1128/MCB.22.21.7417-7427.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Bray GM, Rasminsky M, Aguayo AJ. Interactions between axons and their sheath cells. Annu. Rev. Neurosci. 1981;4:127–162. doi: 10.1146/annurev.ne.04.030181.001015. [DOI] [PubMed] [Google Scholar]
  27. Brittis PA, Canning DR, Silver J. Chondroitin sulfate as a regulator of neuronal patterning in the retina. Science. 1992;255:733–736. doi: 10.1126/science.1738848. [DOI] [PubMed] [Google Scholar]
  28. Brückner G, Brauer K, Hartig W, Wolff JR, Rickmann MJ, Derouiche A, et al. Perineuronal nets provide a polyanionic, glia-associated form of microenvironment around certain neurons in many parts of the rat brain. Glia. 1993;8:183–200. doi: 10.1002/glia.440080306. [DOI] [PubMed] [Google Scholar]
  29. Brückner G, Seeger G, Brauer K, Hartig W, Kacza J, Bigl V. Cortical areas are revealed by distribution patterns of proteoglycan components and parvalbumin in the Mongolian gerbil and rat. Brain Res. 1994;658:67–86. doi: 10.1016/s0006-8993(09)90012-9. [DOI] [PubMed] [Google Scholar]
  30. Brückner G, Bringmann A, Hartig W, Koppe G, Delpech B, Brauer K. Acute and long-lasting changes in extracellular-matrix chondroitin-sulphate proteoglycans induced by injection of chondroitinase ABC in the adult rat brain. Exp. Brain Res. 1998;121:300–310. doi: 10.1007/s002210050463. [DOI] [PubMed] [Google Scholar]
  31. Brückner G, Hausen D, Hartig W, Drlicek M, Arendt T, Brauer K. Cortical areas abundant in extracellular matrix chondroitin sulphate proteoglycans are less affected by cytoskeletal changes in Alzheimer's disease. Neuroscience. 1999;92:791–805. doi: 10.1016/s0306-4522(99)00071-8. [DOI] [PubMed] [Google Scholar]
  32. Brückner G, Grosche J, Schmidt S, Hartig W, Margolis RU, Delpech B, et al. Postnatal development of perineuronal nets in wild-type mice and in a mutant deficient in tenascin-R. J. Comp. Neurol. 2000;428:616–629. doi: 10.1002/1096-9861(20001225)428:4<616::aid-cne3>3.0.co;2-k. [DOI] [PubMed] [Google Scholar]
  33. Bukalo O, Schachner M, Dityatev A. Modification of extracellular matrix by enzymatic removal of chondroitin sulfate and by lack of tenascin-R. differentially affects several forms of synaptic plasticity in the hippocampus. Neuroscience. 2001;104:359–369. doi: 10.1016/s0306-4522(01)00082-3. [DOI] [PubMed] [Google Scholar]
  34. Burg MA, Nishiyama A, Stallcup WB. A central segment of the NG2 proteoglycan is critical for the ability of glioma cells to bind and migrate toward type VI collagen. Exp. Cell Res. 1997;235:254–264. doi: 10.1006/excr.1997.3674. [DOI] [PubMed] [Google Scholar]
  35. Caceres J, Brandan E. Interaction between Alzheimer's disease beta A4 precursor protein (APP) and the extracellular matrix: evidence for the participation of heparan sulfate proteoglycans. J. Cell Biochem. 1997;65:145–158. doi: 10.1002/(sici)1097-4644(199705)65:2<145::aid-jcb2>3.0.co;2-u. [DOI] [PubMed] [Google Scholar]
  36. Canning DR, McKeon RJ, DeWitt DA, Perry G, Wujek JR, Frederickson RC, et al. beta-Amyloid of Alzheimer's disease induces reactive gliosis that inhibits axonal outgrowth. Exp. Neurol. 1993;124:289–298. doi: 10.1006/exnr.1993.1199. [DOI] [PubMed] [Google Scholar]
  37. Celio MR, Blumcke I. Perineuronal nets – a specialized form of extracellular matrix in the adult nervous system. Brain Res. Brain Res. Rev. 1994;19:128–145. doi: 10.1016/0165-0173(94)90006-x. [DOI] [PubMed] [Google Scholar]
  38. Chekenya M, Rooprai HK, Davies D, Levine JM, Butt AM, Pilkington GJ. The NG2 chondroitin sulfate proteoglycan: role in malignant progression of human brain tumours. Int. J. Dev. Neurosci. 1999;17:421–435. doi: 10.1016/s0736-5748(99)00019-2. [DOI] [PubMed] [Google Scholar]
  39. Chen XD, Shi S, Xu T, Robey PG, Young MF. Age-related osteoporosis in biglycan-deficient mice is related to defects in bone marrow stromal cells. J. Bone Miner. Res. 2002;17:331–340. doi: 10.1359/jbmr.2002.17.2.331. [DOI] [PubMed] [Google Scholar]
  40. 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]
  41. Chuong CM, Chen HM. Enhanced expression of neural cell adhesion molecules and tenascin (cytotactin) during wound healing. Am. J. Pathol. 1991;138:427–440. [PMC free article] [PubMed] [Google Scholar]
  42. Crossin KL, Prieto AL, Hoffman S, Jones FS, Friedlander DR. Expression of adhesion molecules and the establishment of boundaries during embryonic and neural development. Exp. Neurol. 1990;109:6–18. doi: 10.1016/s0014-4886(05)80004-4. [DOI] [PubMed] [Google Scholar]
  43. Crutcher KA. Tissue sections from the mature rat brain and spinal cord as substrates for neurite outgrowth in vitro: extensive growth on gray matter but little growth on white matter. Exp. Neurol. 1989;104:39–54. doi: 10.1016/0014-4886(89)90007-1. [DOI] [PubMed] [Google Scholar]
  44. Cynader M, Berman N, Hein A. Recovery of function in cat visual cortex following prolonged deprivation. Exp. Brain Res. 1976;25:139–156. doi: 10.1007/BF00234899. [DOI] [PubMed] [Google Scholar]
  45. David S, Aguayo AJ. Axonal elongation into peripheral nervous system ‘bridges’ after central nervous system injury in adult rats. Science. 1981;214:931–933. doi: 10.1126/science.6171034. [DOI] [PubMed] [Google Scholar]
  46. Davie EW, Fujikawa K, Kisiel W. The coagulation cascade: initiation, maintenance, and regulation. Biochemistry. 1991;30:10363–10370. doi: 10.1021/bi00107a001. [DOI] [PubMed] [Google Scholar]
  47. Davies SJ, Fitch MT, Memberg SP, Hall AK, Raisman G, Silver J. Regeneration of adult axons in white matter tracts of the central nervous system. Nature. 1997;390:680–683. doi: 10.1038/37776. [DOI] [PubMed] [Google Scholar]
  48. Davies SJ, Goucher DR, Doller C, Silver J. Robust regeneration of adult sensory axons in degenerating white matter of the adult rat spinal cord. J. Neurosci. 1999;19:5810–5822. doi: 10.1523/JNEUROSCI.19-14-05810.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Daw NW, Wyatt HJ. Kittens reared in a unidirectional environment: evidence for a critical period. J. Physiol. 1976;257:155–170. doi: 10.1113/jphysiol.1976.sp011361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. De Witt 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]
  51. Dihanich M, Kaser M, Reinhard E, Cunningham D, Monard D. Prothrombin mRNA is expressed by cells of the nervous system. Neuron. 1991;6:575–581. doi: 10.1016/0896-6273(91)90060-d. [DOI] [PubMed] [Google Scholar]
  52. Dityatev A, Schachner M. Extracellular matrix molecules and synaptic plasticity. Nat. Rev. Neurosci. 2003;4:456–468. doi: 10.1038/nrn1115. [DOI] [PubMed] [Google Scholar]
  53. Doege KJ, Sasaki M, Kimura T, Yamada Y. Complete coding sequence and deduced primary structure of the human cartilage large aggregating proteoglycan, aggrecan. Human-specific repeats, and additional alternatively spliced forms. J. Biol. Chem. 1991;266:894–902. [PubMed] [Google Scholar]
  54. Dou CL, Levine JM. Inhibition of neurite growth by the NG2 chondroitin sulfate proteoglycan. J. Neurosci. 1994;14:7616–7628. doi: 10.1523/JNEUROSCI.14-12-07616.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Dou CL, Levine JM. Identification of a neuronal cell surface receptor for a growth inhibitory chondroitin sulfate proteoglycan (NG2) J. Neurochem. 1997;68:1021–1030. doi: 10.1046/j.1471-4159.1997.68031021.x. [DOI] [PubMed] [Google Scholar]
  56. Emerling DE, Lander AD. Inhibitors and promoters of thalamic neuron adhesion and outgrowth in embryonic neocortex: functional association with chondroitin sulfate. Neuron. 1996;17:1089–1100. doi: 10.1016/s0896-6273(00)80242-1. [DOI] [PubMed] [Google Scholar]
  57. Faissner A, Kruse J, Chiquet-Ehrismann R, Mackie E. The high-molecular-weight J1 glycoproteins are immunochemically related to tenascin. Differentiation. 1988;37:104–114. doi: 10.1111/j.1432-0436.1988.tb00802.x. [DOI] [PubMed] [Google Scholar]
  58. Faissner A, Kruse J. J1/tenascin is a repulsive substrate for central nervous system neurons. Neuron. 1990;5:627–637. doi: 10.1016/0896-6273(90)90217-4. [DOI] [PubMed] [Google Scholar]
  59. Faissner A. The tenascin gene family in axon growth and guidance. Cell Tissue Res. 1997;290:331–341. doi: 10.1007/s004410050938. [DOI] [PubMed] [Google Scholar]
  60. Fawcett JW, Housden E, Smith-Thomas L, Meyer RL. The growth of axons in three-dimensional astrocyte cultures. Dev. Biol. 1989a;135:449–458. doi: 10.1016/0012-1606(89)90193-0. [DOI] [PubMed] [Google Scholar]
  61. Fawcett JW, Rokos J, Bakst I. Oligodendrocytes repel axons and cause axonal growth cone collapse. J. Cell Sci. 1989b;92:93–100. doi: 10.1242/jcs.92.1.93. [DOI] [PubMed] [Google Scholar]
  62. Fitch MT, Silver J. Activated macrophages and the blood–brain barrier: inflammation after CNS injury leads to increases in putative inhibitory molecules. Exp. Neurol. 1997;148:587–603. doi: 10.1006/exnr.1997.6701. [DOI] [PubMed] [Google Scholar]
  63. Fitch MT, Doller C, Combs CK, Landreth GE, Silver J. Cellular and molecular mechanisms of glial scarring and progressive cavitation: in vivo and in vitro analysis of inflammation-induced secondary injury after CNS trauma. J. Neurosci. 1999;19:8182–8198. doi: 10.1523/JNEUROSCI.19-19-08182.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Friedlander DR, Milev P, Karthikeyan L, Margolis RK, Margolis RU, Grumet M. The neuronal chondroitin sulfate proteoglycan neurocan binds to the neural cell adhesion molecules Ng-CAM/L1/NILE and N-CAM, and inhibits neuronal adhesion and neurite outgrowth. J. Cell Biol. 1994;125:669–680. doi: 10.1083/jcb.125.3.669. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Gary SC, Zerillo CA, Chiang VL, Gaw JU, Gray G, Hockfield S. cDNA cloning, chromosomal localization, and expression analysis of human BEHAB/brevican, a brain specific proteoglycan regulated during cortical development and in glioma. Gene. 2000;256:139–147. doi: 10.1016/s0378-1119(00)00362-0. [DOI] [PubMed] [Google Scholar]
  66. Gingrich MB, Traynelis SF. Serine proteases and brain damage – is there a link? Trends Neurosci. 2000;23:399–407. doi: 10.1016/s0166-2236(00)01617-9. [DOI] [PubMed] [Google Scholar]
  67. Giulian D, Lachman LB. Interleukin-1 stimulation of astroglial proliferation after brain injury. Science. 1985;228:497–499. doi: 10.1126/science.3872478. [DOI] [PubMed] [Google Scholar]
  68. Giulian D, Woodward J, Young DG, Krebs JF, Lachman LB. Interleukin-1 injected into mammalian brain stimulates astrogliosis and neovascularization. J. Neurosci. 1988;8:2485–2490. doi: 10.1523/JNEUROSCI.08-07-02485.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Glumoff V, Savontaus M, Vehanen J, Vuorio E. Analysis of aggrecan and tenascin gene expression in mouse skeletal tissues by northern and in situ hybridization using species specific cDNA probes. Biochim. Biophys. Acta. 1994;1219:613–622. doi: 10.1016/0167-4781(94)90220-8. [DOI] [PubMed] [Google Scholar]
  70. Goedert M, Jakes R, Spillantini MG, Hasegawa M, Smith MJ, Crowther RA. Assembly of microtubule-associated protein tau into Alzheimer-like filaments induced by sulphated glycosaminoglycans. Nature. 1996;383:550–553. doi: 10.1038/383550a0. [DOI] [PubMed] [Google Scholar]
  71. Grabham P, Cunningham DD. Thrombin receptor activation stimulates astrocyte proliferation and reversal of stellation by distinct pathways: involvement of tyrosine phosphorylation. J. Neurochem. 1995;64:583–591. doi: 10.1046/j.1471-4159.1995.64020583.x. [DOI] [PubMed] [Google Scholar]
  72. Grand RJ, Turnell AS, Grabham PW. Cellular consequences of thrombin-receptor activation. Biochem. J. 1996;313:353–368. doi: 10.1042/bj3130353. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Grumet M, Flaccus A, Margolis RU. Functional characterization of chondroitin sulfate proteoglycans of brain: interactions with neurons and neural cell adhesion molecules. J. Cell Biol. 1993;120:815–824. doi: 10.1083/jcb.120.3.815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Grumet M, Milev P, Sakurai T, Karthikeyan L, Bourdon M, Margolis RK, et al. Interactions with tenascin and differential effects on cell adhesion of neurocan and phosphacan, two major chondroitin sulfate proteoglycans of nervous tissue. J. Biol. Chem. 1994;269:12142–12146. [PubMed] [Google Scholar]
  75. Guimaraes A, Zaremba S, Hockfield S. Molecular and morphological changes in the cat lateral geniculate nucleus and visual cortex induced by visual deprivation are revealed by monoclonal antibodies Cat-304 and Cat-301. J. Neurosci. 1990;10:3014–3024. doi: 10.1523/JNEUROSCI.10-09-03014.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Gurwitz D, Cunningham DD. Thrombin modulates and reverses neuroblastoma neurite outgrowth. Proc. Natl Acad. Sci. USA. 1988;85:3440–3444. doi: 10.1073/pnas.85.10.3440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Haas CA, Rauch U, Thon N, Merten T, Deller T. Entorhinal cortex lesion in adult rats induces the expression of the neuronal chondroitin sulfate proteoglycan neurocan in reactive astrocytes. J. Neurosci. 1999;19:9953–9963. doi: 10.1523/JNEUROSCI.19-22-09953.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Hartig W, Brauer K, Bruckner G. Wisteria floribunda agglutinin-labelled nets surround parvalbumin-containing neurons. Neuroreport. 1992;3:869–872. doi: 10.1097/00001756-199210000-00012. [DOI] [PubMed] [Google Scholar]
  79. Hartig W, Brauer K, Bigl V, Bruckner G. Chondroitin sulfate proteoglycan-immunoreactivity of lectin-labeled perineuronal nets around parvalbumin-containing neurons. Brain Res. 1994;635:307–311. doi: 10.1016/0006-8993(94)91452-4. [DOI] [PubMed] [Google Scholar]
  80. Hartig W, Klein C, Brauer K, Schuppel KF, Arendt T, Bigl V, et al. Hyperphosphorylated protein tau is restricted to neurons devoid of perineuronal nets in the cortex of aged bison. Neurobiol. Aging. 2001;22:25–33. doi: 10.1016/s0197-4580(00)00179-2. [DOI] [PubMed] [Google Scholar]
  81. Haunso A, Celio MR, Margolis RK, Menoud PA. Phosphacan immunoreactivity is associated with perineuronal nets around parvalbumin-expressing neurones. Brain Res. 1999;834:219–222. doi: 10.1016/s0006-8993(99)01596-6. [DOI] [PubMed] [Google Scholar]
  82. Haunso A, Ibrahim M, Bartsch U, Letiembre M, Celio MR, Menoud P. Morphology of perineuronal nets in tenascin-R. and parvalbumin single and double knockout mice. Brain Res. 2000;864:142–145. doi: 10.1016/s0006-8993(00)02173-9. [DOI] [PubMed] [Google Scholar]
  83. Hausen D, Bruckner G, Drlicek M, Hartig W, Brauer K, Bigl V. Pyramidal cells ensheathed by perineuronal nets in human motor and somatosensory cortex. Neuroreport. 1996;7:1725–1729. doi: 10.1097/00001756-199607290-00006. [DOI] [PubMed] [Google Scholar]
  84. Hedbom E, Heinegard D. Binding of fibromodulin and decorin to separate sites on fibrillar collagens. J. Biol. Chem. 1993;268:27307–27312. [PubMed] [Google Scholar]
  85. Hockfield S, Kalb RG, Zaremba S, Fryer H. Expression of neural proteoglycans correlates with the acquisition of mature neuronal properties in the mammalian brain. Cold Spring Harb. Symp. Quant. Biol. 1990;55:505–514. doi: 10.1101/sqb.1990.055.01.049. [DOI] [PubMed] [Google Scholar]
  86. Hockfield S, Kalb RG. Activity-dependent structural changes during neuronal development. Curr. Opin. Neurobiol. 1993;3:87–92. doi: 10.1016/0959-4388(93)90040-6. [DOI] [PubMed] [Google Scholar]
  87. Hoffman S, Crossin KL, Edelman GM. Molecular forms, binding functions, and developmental expression patterns of cytotactin and cytotactin-binding proteoglycan, an interactive pair of extracellular matrix molecules. J. Cell Biol. 1988;106:519–532. doi: 10.1083/jcb.106.2.519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Husmann K, Carbonetto S, Schachner M. Distinct sites on tenascin-C mediate repellent or adhesive interactions with different neuronal cell types. Cell Adhes. Commun. 1995;3:293–310. doi: 10.3109/15419069509081015. [DOI] [PubMed] [Google Scholar]
  89. Jepsen KJ, Wu F, Peragallo JH, Paul J, Roberts L, Ezura Y, et al. A syndrome of joint laxity and impaired tendon integrity in lumican- and fibromodulin-deficient mice. J. Biol. Chem. 2002;277:35532–35540. doi: 10.1074/jbc.M205398200. [DOI] [PubMed] [Google Scholar]
  90. Jones LL, Margolis RU, Tuszynski MH. The chondroitin sulfate proteoglycans neurocan, brevican, phosphacan, and versican are differentially regulated following spinal cord injury. Exp. Neurol. 2003;182:399–411. doi: 10.1016/s0014-4886(03)00087-6. [DOI] [PubMed] [Google Scholar]
  91. Kahn MA, Ellison JA, Speight GJ, de Vellis J. CNTF regulation of astrogliosis and the activation of microglia in the developing rat central nervous system. Brain Res. 1995;685:55–67. doi: 10.1016/0006-8993(95)00411-i. [DOI] [PubMed] [Google Scholar]
  92. Kalb RG, Hockfield S. Molecular evidence for early activity-dependent development of hamster motor neurons. J. Neurosci. 1988;8:2350–2360. doi: 10.1523/JNEUROSCI.08-07-02350.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Kalb RG, Hockfield S. Induction of a neuronal proteoglycan by the NMDA receptor in the developing spinal cord. Science. 1990;250:294–296. doi: 10.1126/science.2145629. [DOI] [PubMed] [Google Scholar]
  94. Kalil K, Reh T. A light and electron microscopic study of regrowing pyramidal tract fibers. J. Comp Neurol. 1982;211:265–275. doi: 10.1002/cne.902110305. [DOI] [PubMed] [Google Scholar]
  95. Knudson W, Knudson CB. Assembly of a chondrocyte-like pericellular matrix on non-chondrogenic cells. Role of the cell surface hyaluronan receptors in the assembly of a pericellular matrix. J. Cell Sci. 1991;99:227–235. doi: 10.1242/jcs.99.2.227. [DOI] [PubMed] [Google Scholar]
  96. Koppe G, Bruckner G, Brauer K, Hartig W, Bigl V. Developmental patterns of proteoglycan-containing extracellular matrix in perineuronal nets and neuropil of the postnatal rat brain. Cell Tissue Res. 1997a;288:33–41. doi: 10.1007/s004410050790. [DOI] [PubMed] [Google Scholar]
  97. Koppe G, Bruckner G, Hartig W, Delpech B, Bigl V. Characterization of proteoglycan-containing perineuronal nets by enzymatic treatments of rat brain sections. Histochem. J. 1997b;29:11–20. doi: 10.1023/a:1026408716522. [DOI] [PubMed] [Google Scholar]
  98. Krekoski CA, Neubauer D, Zuo J, Muir D. Axonal regeneration into acellular nerve grafts is enhanced by degradation of chondroitin sulfate proteoglycan. J. Neurosci. 2001;21:6206–6213. doi: 10.1523/JNEUROSCI.21-16-06206.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Krusius T, Gehlsen KR, Ruoslahti E. A fibroblast chondroitin sulfate proteoglycan core protein contains lectin-like and growth factor-like sequences. J. Biol. Chem. 1987;262:13120–13125. [PubMed] [Google Scholar]
  100. Laywell ED, Steindler DA. Boundaries and wounds, glia and glycoconjugates. Cellular and molecular analyses of developmental partitions and adult brain lesions. Ann. NY Acad. Sci. 1991;633:122–141. doi: 10.1111/j.1749-6632.1991.tb15603.x. [DOI] [PubMed] [Google Scholar]
  101. Laywell ED, Dorries U, Bartsch U, Faissner A, Schachner M, Steindler DA. Enhanced expression of the developmentally regulated extracellular matrix molecule tenascin following adult brain injury. Proc. Natl Acad. Sci. USA. 1992;89:2634–2638. doi: 10.1073/pnas.89.7.2634. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Levine JM, Card JP. Light and electron microscopic localization of a cell surface antigen (NG2) in the rat cerebellum: association with smooth protoplasmic astrocytes. J. Neurosci. 1987;7:2711–2720. doi: 10.1523/JNEUROSCI.07-09-02711.1987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Levine JM, Stincone F, Lee YS. Development and differentiation of glial precursor cells in the rat cerebellum. Glia. 1993;7:307–321. doi: 10.1002/glia.440070406. [DOI] [PubMed] [Google Scholar]
  104. Levine JM. Increased expression of the NG2 chondroitin-sulfate proteoglycan after brain injury. J. Neurosci. 1994;14:4716–4730. doi: 10.1523/JNEUROSCI.14-08-04716.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Liang Y, Haring M, Roughley PJ, Margolis RK, Margolis RU. Glypican and biglycan in the nuclei of neurons and glioma cells: presence of functional nuclear localization signals and dynamic changes in glypican during the cell cycle. J. Cell Biol. 1997;139:851–864. doi: 10.1083/jcb.139.4.851. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Lochter A, Vaughan L, Kaplony A, Prochiantz A, Schachner M, Faissner A. J1/tenascin in substrate-bound and soluble form displays contrary effects on neurite outgrowth. J. Cell Biol. 1991;113:1159–1171. doi: 10.1083/jcb.113.5.1159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Maeda N, Hamanaka H, Shintani T, Nishiwaki T, Noda M. Multiple receptor-like protein tyrosine phosphatases in the form of chondroitin sulfate proteoglycan. FEBS Lett. 1994;354:67–70. doi: 10.1016/0014-5793(94)01093-5. [DOI] [PubMed] [Google Scholar]
  108. Maleski M, Hockfield S. Glial cells assemble hyaluronan-based pericellular matrices in vitro. Glia. 1997;20:193–202. doi: 10.1002/(sici)1098-1136(199707)20:3<193::aid-glia3>3.0.co;2-9. [DOI] [PubMed] [Google Scholar]
  109. Margolis RK, Margolis RU. Nervous tissue proteoglycans. Experientia. 1993;49:429–446. doi: 10.1007/BF01923587. [DOI] [PubMed] [Google Scholar]
  110. Margolis RU, Margolis RK. Chondroitin sulfate proteoglycans as mediators of axon growth and pathfinding. Cell Tissue Res. 1997;290:343–348. doi: 10.1007/s004410050939. [DOI] [PubMed] [Google Scholar]
  111. Matsui F, Nishizuka M, Yasuda Y, Aono S, Watanabe E, Oohira A. Occurrence of a N-terminal proteolytic fragment of neurocan, not a C-terminal half, in a perineuronal net in the adult rat cerebrum. Brain Res. 1998;790:45–51. doi: 10.1016/s0006-8993(98)00009-2. [DOI] [PubMed] [Google Scholar]
  112. Matthews RT, Kelly GM, Zerillo CA, Gray G, Tiemeyer M, Hockfield S. Aggrecan glycoforms contribute to the molecular heterogeneity of perineuronal nets. J. Neurosci. 2002;22:7536–7547. doi: 10.1523/JNEUROSCI.22-17-07536.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. 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]
  114. Maxwell WL, Follows R, Ashhurst DE, Berry M. The response of the cerebral hemisphere of the rat to injury. I. The mature rat. Philos. Trans. R. Soc. Lond. B Biol. Sci. 1990;328:479–500. doi: 10.1098/rstb.1990.0121. [DOI] [PubMed] [Google Scholar]
  115. McGee AW, Strittmatter SM. The Nogo-66 receptor: focusing myelin inhibition of axon regeneration. Trends Neurosci. 2003;26:193–198. doi: 10.1016/S0166-2236(03)00062-6. [DOI] [PubMed] [Google Scholar]
  116. McKeon RJ, Schreiber RC, Rudge JS, Silver J. Reduction of neurite outgrowth in a model of glial scarring following CNS injury is correlated with the expression of inhibitory molecules on reactive astrocytes. J. Neurosci. 1991;11:3398–3411. doi: 10.1523/JNEUROSCI.11-11-03398.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. McKeon RJ, Hoke A, Silver J. Injury-induced proteoglycans inhibit the potential for laminin-mediated axon growth on astrocytic scars. Exp. Neurol. 1995;136:32–43. doi: 10.1006/exnr.1995.1081. [DOI] [PubMed] [Google Scholar]
  118. McKeon RJ, Jurynec MJ, Buck CR. The chondroitin sulfate proteoglycans neurocan and phosphacan are expressed by reactive astrocytes in the chronic CNS glial scar. J. Neurosci. 1999;19:10778–10788. doi: 10.1523/JNEUROSCI.19-24-10778.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Meiners S, Mercado ML, Nure-Kamal MS, Geller HM. Tenascin-C contains domains that independently regulate neurite outgrowth and neurite guidance. J. Neurosci. 1999;19:8443–8453. doi: 10.1523/JNEUROSCI.19-19-08443.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Meiners S, Nure-Kamal MS, Mercado ML. Identification of a neurite outgrowth-promoting motif within the alternatively spliced region of human tenascin-C. J. Neurosci. 2001;21:7215–7225. doi: 10.1523/JNEUROSCI.21-18-07215.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Meyer-Puttlitz B, Junker E, Margolis RU, Margolis RK. Chondroitin sulfate proteoglycans in the developing central nervous system. II. Immunocytochemical localization of neurocan and phosphacan. J. Comp. Neurol. 1996;366:44–54. doi: 10.1002/(SICI)1096-9861(19960226)366:1<44::AID-CNE4>3.0.CO;2-K. [DOI] [PubMed] [Google Scholar]
  122. Milev P, Friedlander DR, Sakurai T, Karthikeyan L, Flad M, Margolis RK, et al. Interactions of the chondroitin sulfate proteoglycan phosphacan, the extracellular domain of a receptor-type protein tyrosine phosphatase, with neurons, glia, and neural cell adhesion molecules. J. Cell Biol. 1994;127:1703–1715. doi: 10.1083/jcb.127.6.1703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Milev P, Maurel P, Chiba A, Mevissen M, Popp S, Yamaguchi Y, et al. Differential regulation of expression of hyaluronan-binding proteoglycans in developing brain: aggrecan, versican, neurocan, and brevican. Biochem. Biophys. Res. Commun. 1998;247:207–212. doi: 10.1006/bbrc.1998.8759. [DOI] [PubMed] [Google Scholar]
  124. Möller T, Hanisch UK, Ransom BR. Thrombin-induced activation of cultured rodent microglia. J. Neurochem. 2000;75:1539–1547. doi: 10.1046/j.1471-4159.2000.0751539.x. [DOI] [PubMed] [Google Scholar]
  125. Moon LD, Asher RA, Rhodes KE, Fawcett JW. Regeneration of CNS axons back to their target following treatment of adult rat brain with chondroitinase ABC. Nat. Neurosci. 2001;4:465–466. doi: 10.1038/87415. [DOI] [PubMed] [Google Scholar]
  126. 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]
  127. Mundlos S, Meyer R, Yamada Y, Zabel B. Distribution of cartilage proteoglycan (aggrecan) core protein and link protein gene expression during human skeletal development. Matrix. 1991;11:339–346. doi: 10.1016/s0934-8832(11)80205-2. [DOI] [PubMed] [Google Scholar]
  128. Murakami T, Ohtsuka A, Taguchi T, Piao DX. Perineuronal sulfated proteoglycans and dark neurons in the brain and spinal cord: a histochemical and electron microscopic study of newborn and adult mice. Arch. Histol. Cytol. 1995;58:557–565. doi: 10.1679/aohc.58.557. [DOI] [PubMed] [Google Scholar]
  129. Nerucci F, Fioravanti A, Cicero MR, Collodel G, Marcolongo R. Effects of chondroitin sulfate and interleukin-1beta on human chondrocyte cultures exposed to pressurization: a biochemical and morphological study. Osteoarthritis Cartilage. 2000;8:279–287. doi: 10.1053/joca.1999.0302. [DOI] [PubMed] [Google Scholar]
  130. Niclou SP, Suidan HS, Pavlik A, Vejsada R, Monard D. Changes in the expression of protease-activated receptor 1 and protease nexin-1 mRNA during rat nervous system development and after nerve lesion. Eur. J. Neurosci. 1998;10:1590–1607. doi: 10.1046/j.1460-9568.1998.00183.x. [DOI] [PubMed] [Google Scholar]
  131. Niederöst BP, Zimmermann DR, Schwab ME, Bandtlow CE. Bovine CNS myelin contains neurite growth-inhibitory activity associated with chondroitin sulfate proteoglycans. J. Neurosci. 1999;19:8979–8989. doi: 10.1523/JNEUROSCI.19-20-08979.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Nishino A, Suzuki M, Ohtani H, Motohashi O, Umezawa K, Nagura H, et al. Thrombin may contribute to the pathophysiology of central nervous system injury. J. Neurotrauma. 1993;10:167–179. doi: 10.1089/neu.1993.10.167. [DOI] [PubMed] [Google Scholar]
  133. Nishiyama A, Lin XH, Stallcup WB. Generation of truncated forms of the NG2 proteoglycan by cell surface proteolysis. Mol. Biol. Cell. 1995;6:1819–1832. doi: 10.1091/mbc.6.12.1819. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Nutt CL, Matthews RT, Hockfield S. Glial tumor invasion: a role for the upregulation and cleavage of BEHAB/brevican. Neuroscientist. 2001;7:113–122. doi: 10.1177/107385840100700206. [DOI] [PubMed] [Google Scholar]
  135. Oakley RA, Tosney KW. Peanut agglutinin and chondroitin-6-sulfate are molecular markers for tissues that act as barriers to axon advance in the avian embryo. Dev. Biol. 1991;147:187–206. doi: 10.1016/s0012-1606(05)80017-x. [DOI] [PubMed] [Google Scholar]
  136. Okamoto M, Mori S, Endo H. A protective action of chondroitin sulfate proteoglycans against neuronal cell death induced by glutamate. Brain Res. 1994a;637:57–67. doi: 10.1016/0006-8993(94)91217-3. [DOI] [PubMed] [Google Scholar]
  137. Okamoto M, Mori S, Ichimura M, Endo H. Chondroitin sulfate proteoglycans protect cultured rat's cortical and hippocampal neurons from delayed cell death induced by excitatory amino acids. Neurosci. Lett. 1994b;172:51–54. doi: 10.1016/0304-3940(94)90660-2. [DOI] [PubMed] [Google Scholar]
  138. Ong WY, Levine JM. A light and electron microscopic study of NG2 chondroitin sulfate proteoglycan-positive oligodendrocyte precursor cells in the normal and kainate-lesioned rat hippocampus. Neuroscience. 1999;92:83–95. doi: 10.1016/s0306-4522(98)00751-9. [DOI] [PubMed] [Google Scholar]
  139. Oohira A, Matsui F, Watanabe E, Kushima Y, Maeda N. Developmentally regulated expression of a brain specific species of chondroitin sulfate proteoglycan, neurocan, identified with a monoclonal antibody IG2 in the rat cerebrum. Neuroscience. 1994;60:145–157. doi: 10.1016/0306-4522(94)90210-0. [DOI] [PubMed] [Google Scholar]
  140. Paulsson M, Morgelin M, Wiedemann H, Beardmore-Gray M, Dunham D, Hardingham T, et al. Extended and globular protein domains in cartilage proteoglycans. Biochem. J. 1987;245:763–772. doi: 10.1042/bj2450763. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Paulus W, Baur I, Dours-Zimmermann MT, Zimmermann DR. Differential expression of versican isoforms in brain tumors. J. Neuropathol. Exp. Neurol. 1996;55:528–533. doi: 10.1097/00005072-199605000-00005. [DOI] [PubMed] [Google Scholar]
  142. Perry VH, Brown MC, Gordon S. The macrophage response to central and peripheral nerve injury. A possible role for macrophages in regeneration. J. Exp. Med. 1987;165:1218–1223. doi: 10.1084/jem.165.4.1218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. 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]
  144. Pettigrew JD. The importance of early visual experience for neurons of the developing geniculostriate system. Invest. Ophthalmol. 1972;11:386–394. [PubMed] [Google Scholar]
  145. Pizzorusso T, Medini P, Berardi N, Chierzi S, Fawcett JW, Maffei L. Reactivation of ocular dominance plasticity in the adult visual cortex. Science. 2002;298:1248–1251. doi: 10.1126/science.1072699. [DOI] [PubMed] [Google Scholar]
  146. Popp S, Andersen JS, Maurel P, Margolis RU. Localization of aggrecan and versican in the developing rat central nervous system. Dev. Dyn. 2003;227:143–149. doi: 10.1002/dvdy.10282. [DOI] [PubMed] [Google Scholar]
  147. Pulkkinen L, Kainulainen K, Krusius T, Makinen P, Schollin J, Gustavsson KH, et al. Deficient expression of the gene coding for decorin in a lethal form of Marfan syndrome. J. Biol. Chem. 1990;265:17780–17785. [PubMed] [Google Scholar]
  148. Raivich G, Bohatschek M, Kloss CU, Werner A, Jones LL, Kreutzberg GW. Neuroglial activation repertoire in the injured brain: graded response, molecular mechanisms and cues to physiological function. Brain Res. Brain Res. Rev. 1999;30:77–105. doi: 10.1016/s0165-0173(99)00007-7. [DOI] [PubMed] [Google Scholar]
  149. Rauch U, Karthikeyan L, Maurel P, Margolis RU, Margolis RK. Cloning and primary structure of neurocan, a developmentally regulated, aggregating chondroitin sulfate proteoglycan of brain. J. Biol. Chem. 1992;267:19536–19547. [PubMed] [Google Scholar]
  150. Rhodes KE, Raivich G, Fawcett JW. Regulation of oligodendrocyte precursor cell reactivity in the rat central nervous system. Soc. Neurosci. Abstract. 2002;32:891.10. [Google Scholar]
  151. Rhodes KE, Moon LD, Fawcett JW. Inhibiting cell proliferation during formation of the glial scar: effects on axon regeneration in the CNS. Neuroscience. 2003;120:41–56. doi: 10.1016/s0306-4522(03)00285-9. [DOI] [PubMed] [Google Scholar]
  152. Richardson PM, McGuinness UM, Aguayo AJ. Axons from CNS neurons regenerate into PNS grafts. Nature. 1980;284:264–265. doi: 10.1038/284264a0. [DOI] [PubMed] [Google Scholar]
  153. Richardson PM, Issa VM, Aguayo AJ. Regeneration of long spinal axons in the rat. J. Neurocytol. 1984;13:165–182. doi: 10.1007/BF01148324. [DOI] [PubMed] [Google Scholar]
  154. Richardson RR. Congenital genetic murine (ch) hydrocephalus. A structural model of cellular dysplasia and disorganization with the molecular locus of deficient proteoglycan synthesis. Childs Nerv. Syst. 1985;1:87–99. doi: 10.1007/BF00706688. [DOI] [PubMed] [Google Scholar]
  155. Ridet JL, Malhotra SK, Privat A, Gage FH. Reactive astrocytes: cellular and molecular cues to biological function. Trends Neurosci. 1997;20:570–577. doi: 10.1016/s0166-2236(97)01139-9. [DOI] [PubMed] [Google Scholar]
  156. Rudge JS, Smith GM, Silver J. An in vitro model of wound healing in the CNS: analysis of cell reaction and interaction at different ages. Exp. Neurol. 1989;103:1–16. doi: 10.1016/0014-4886(89)90180-5. [DOI] [PubMed] [Google Scholar]
  157. Savio T, Schwab ME. Rat CNS white matter, but not gray matter, is nonpermissive for neuronal cell adhesion and fiber outgrowth. J. Neurosci. 1989;9:1126–1133. doi: 10.1523/JNEUROSCI.09-04-01126.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Schmalfeldt M, Bandtlow CE, Dours-Zimmermann MT, Winterhalter KH, Zimmermann DR. Brain derived versican V2 is a potent inhibitor of axonal growth. J. Cell Sci. 2000;113:807–816. doi: 10.1242/jcs.113.5.807. [DOI] [PubMed] [Google Scholar]
  159. Schmidt G, Hausser H, Kresse H. Interaction of the small proteoglycan decorin with fibronectin. Involvement of the sequence NKISK of the core protein. Biochem. J. 1991;280:411–414. doi: 10.1042/bj2800411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Schrappe M, Klier FG, Spiro RC, Waltz TA, Reisfeld RA, Gladson CL. Correlation of chondroitin sulfate proteoglycan expression on proliferating brain capillary endothelial cells with the malignant phenotype of astroglial cells. Cancer Res. 1991;51:4986–4993. [PubMed] [Google Scholar]
  161. Schreyer DJ, Jones EG. Growing corticospinal axons by-pass lesions of neonatal rat spinal cord. Neuroscience. 1983;9:31–40. doi: 10.1016/0306-4522(83)90044-1. [DOI] [PubMed] [Google Scholar]
  162. Schubert D, Schroeder R, LaCorbiere M, Saitoh T, Cole G. Amyloid beta protein precursor is possibly a heparan sulfate proteoglycan core protein. Science. 1988;241:223–226. doi: 10.1126/science.2968652. [DOI] [PubMed] [Google Scholar]
  163. Schwab ME, Thoenen H. Dissociated neurons regenerate into sciatic but not optic nerve explants in culture irrespective of neurotrophic factors. J. Neurosci. 1985;5:2415–2423. doi: 10.1523/JNEUROSCI.05-09-02415.1985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Schwab ME, Caroni P. Oligodendrocytes and CNS myelin are nonpermissive substrates for neurite growth and fibroblast spreading in vitro. J. Neurosci. 1988;8:2381–2393. doi: 10.1523/JNEUROSCI.08-07-02381.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Schwartz NB, Domowicz M. Chondrodysplasias due to proteoglycan defects. Glycobiology. 2002;12:57R–68R. doi: 10.1093/glycob/12.4.57r. [DOI] [PubMed] [Google Scholar]
  166. Seidenbecher CI, Richter K, Rauch U, Fassler R, Garner CC, Gundelfinger ED. Brevican, a chondroitin sulfate proteoglycan of rat brain, occurs as secreted and cell surface glycosylphosphatidylinositol-anchored isoforms. J. Biol. Chem. 1995;270:27206–27212. doi: 10.1074/jbc.270.45.27206. [DOI] [PubMed] [Google Scholar]
  167. Serpinskaya AS, Feng G, Sanes JR, Craig AM. Synapse formation by hippocampal neurons from agrin-deficient mice. Dev. Biol. 1999;205:65–78. doi: 10.1006/dbio.1998.9112. [DOI] [PubMed] [Google Scholar]
  168. Shioi J, Anderson JP, Ripellino JA, Robakis NK. Chondroitin sulfate proteoglycan form of the Alzheimer's beta-amyloid precursor. J. Biol. Chem. 1992;267:13819–13822. [PubMed] [Google Scholar]
  169. Shioi J, Refolo LM, Efthimiopoulos S, Robakis NK. Chondroitin sulfate proteoglycan form of cellular and cell-surface Alzheimer amyloid precursor. Neurosci. Lett. 1993;154:121–124. doi: 10.1016/0304-3940(93)90186-o. [DOI] [PubMed] [Google Scholar]
  170. Shoshan Y, Nishiyama A, Chang A, Mork S, Barnett GH, Cowell JK, et al. Expression of oligodendrocyte progenitor cell antigens by gliomas: implications for the histogenesis of brain tumors. Proc. Natl Acad. Sci. USA. 1999;96:10361–10366. doi: 10.1073/pnas.96.18.10361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Smith-Thomas LC, Fok-Seang J, Du Stevens JJS, Muir E, Faissner A, Geller HM, et al. An inhibitor of neurite outgrowth produced by astrocytes. J. Cell Sci. 1994;107:1687–1695. doi: 10.1242/jcs.107.6.1687. [DOI] [PubMed] [Google Scholar]
  172. Smith-Thomas LC, Stevens J, Fok-Seang J, Faissner A, Rogers JH, Fawcett JW. Increased axon regeneration in astrocytes grown in the presence of proteoglycan synthesis inhibitors. J. Cell Sci. 1995;108:1307–1315. doi: 10.1242/jcs.108.3.1307. [DOI] [PubMed] [Google Scholar]
  173. Snow DM, Lemmon V, Carrino DA, Caplan AI, Silver J. Sulfated proteoglycans in astroglial barriers inhibit neurite outgrowth in vitro. Exp. Neurol. 1990a;109:111–130. doi: 10.1016/s0014-4886(05)80013-5. [DOI] [PubMed] [Google Scholar]
  174. Snow AD, Mar H, Nochlin D, Kimata K, Kato M, Suzuki S, et al. 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]
  175. Snow DM, Steindler DA, Silver J. Molecular and cellular characterization of the glial roof plate of the spinal cord and optic tectum: a possible role for a proteoglycan in the development of an axon barrier. Dev. Biol. 1990b;138:359–376. doi: 10.1016/0012-1606(90)90203-u. [DOI] [PubMed] [Google Scholar]
  176. 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]
  177. Stallcup WB, Beasley L. Bipotential glial precursor cells of the optic nerve express the NG2 proteoglycan. J. Neurosci. 1987;7:2737–2744. doi: 10.1523/JNEUROSCI.07-09-02737.1987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Steck PA, Moser RP, Bruner JM, Liang L, Freidman AN, Hwang TL, et al. Altered expression and distribution of heparan sulfate proteoglycans in human gliomas. Cancer Res. 1989;49:2096–2103. [PubMed] [Google Scholar]
  179. Steindler DA, Cooper NG, Faissner A, Schachner M. Boundaries defined by adhesion molecules during development of the cerebral cortex: the J1/tenascin glycoprotein in the mouse somatosensory cortical barrel field. Dev. Biol. 1989;131:243–260. doi: 10.1016/s0012-1606(89)80056-9. [DOI] [PubMed] [Google Scholar]
  180. Stern CD, Norris WE, Bronner-Fraser M, Carlson GJ, Faissner A, Keynes RJ, et al. J1/tenascin-related molecules are not responsible for the segmented pattern of neural crest cells or motor axons in the chick embryo. Development. 1989;107:309–319. doi: 10.1242/dev.107.2.309. [DOI] [PubMed] [Google Scholar]
  181. Stoll G, Trapp BD, Griffin JW. Macrophage function during Wallerian degeneration of rat optic nerve: clearance of degenerating myelin and Ia expression. J. Neurosci. 1989;9:2327–2335. doi: 10.1523/JNEUROSCI.09-07-02327.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Streit WJ, Kincaid-Colton CA. The brain's immune system. Sci. Am. 1995;273:54–61. doi: 10.1038/scientificamerican1195-54. [DOI] [PubMed] [Google Scholar]
  183. Su JH, Cummings BJ, Cotman CW. Localization of heparan sulfate glycosaminoglycan and proteoglycan core protein in aged brain and Alzheimer's disease. Neuroscience. 1992;51:801–813. doi: 10.1016/0306-4522(92)90521-3. [DOI] [PubMed] [Google Scholar]
  184. Sur M, Frost DO, Hockfield S. Expression of a surface-associated antigen on Y-cells in the cat lateral geniculate nucleus is regulated by visual experience. J. Neurosci. 1988;8:874–882. doi: 10.1523/JNEUROSCI.08-03-00874.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Takahashi-Iwanaga H, Murakami T, Abe K. Three-dimensional microanatomy of perineuronal proteoglycan nets enveloping motor neurons in the rat spinal cord. J. Neurocytol. 1998;27:817–827. doi: 10.1023/a:1006955414939. [DOI] [PubMed] [Google Scholar]
  186. Tang X, Davies JE, Davies SJ. 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]
  187. Touab M, Villena J, Barranco C, Arumi-Uria M, Bassols A. Versican is differentially expressed in human melanoma and may play a role in tumor development. Am. J. Pathol. 2002;160:549–557. doi: 10.1016/S0002-9440(10)64874-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Treloar HB, Nurcombe V, Key B. Expression of extracellular matrix molecules in the embryonic rat olfactory pathway. J. Neurobiol. 1996;31:41–55. doi: 10.1002/(SICI)1097-4695(199609)31:1<41::AID-NEU4>3.0.CO;2-F. [DOI] [PubMed] [Google Scholar]
  189. Tropea D, Caleo M, Maffei L. Synergistic effects of brain-derived neurotrophic factor and chondroitinase ABC on retinal fiber sprouting after denervation of the superior colliculus in adult rats. J. Neurosci. 2003;23:7034–7044. doi: 10.1523/JNEUROSCI.23-18-07034.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Uebelhart D, Thonar EJ, Zhang J, Williams JM. Protective effect of exogenous chondroitin 4,6-sulfate in the acute degradation of articular cartilage in the rabbit. Osteoarthritis Cartilage. 1998;6(Suppl. A):6–13. doi: 10.1016/s1063-4584(98)80005-8. [DOI] [PubMed] [Google Scholar]
  191. Watanabe E, Aono S, Matsui F, Yamada Y, Naruse I, Oohira A. Distribution of a brain-specific proteoglycan, neurocan, and the corresponding mRNA during the formation of barrels in the rat somatosensory cortex. Eur. J. Neurosci. 1995;7:547–554. doi: 10.1111/j.1460-9568.1995.tb00659.x. [DOI] [PubMed] [Google Scholar]
  192. Weber P, Bartsch U, Rasband MN, Czaniera R, Lang Y, Bluethmann H, et al. Mice deficient for tenascin-R. display alterations of the extracellular matrix and decreased axonal conduction velocities in the CNS. J. Neurosci. 1999;19:4245–4262. doi: 10.1523/JNEUROSCI.19-11-04245.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Weinstein JR, Gold SJ, Cunningham DD, Gall CM. Cellular localization of thrombin receptor mRNA in rat brain: expression by mesencephalic dopaminergic neurons and codistribution with prothrombin mRNA. J. Neurosci. 1995;15:2906–2919. doi: 10.1523/JNEUROSCI.15-04-02906.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Winter CG, Saotome Y, Levison SW, Hirsh D. A role for ciliary neurotrophic factor as an inducer of reactive gliosis, the glial response to central nervous system injury. Proc. Natl Acad. Sci. USA. 1995;92:5865–5869. doi: 10.1073/pnas.92.13.5865. [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Wintergerst ES, Faissner A, Celio MR. The proteoglycan DSD-1-PG occurs in perineuronal nets around parvalbumin-immunoreactive interneurons of the rat cerebral cortex. Int. J. Dev. Neurosci. 1996;14:249–255. doi: 10.1016/0736-5748(96)00011-1. [DOI] [PubMed] [Google Scholar]
  196. Woods AG, Cribbs DH, Whittemore ER, Cotman CW. Heparan sulfate and chondroitin sulfate glycosaminoglycan attenuate beta-amyloid (25–35) induced neurodegeneration in cultured hippocampal neurons. Brain Res. 1995;697:53–62. doi: 10.1016/0006-8993(95)00775-l. [DOI] [PubMed] [Google Scholar]
  197. Xiao ZC, Bartsch U, Margolis RK, Rougon G, Montag D, Schachner M. Isolation of a tenascin-R. binding protein from mouse brain membranes. A phosphacan-related chondroitin sulfate proteoglycan. J. Biol. Chem. 1997;272:32092–32101. doi: 10.1074/jbc.272.51.32092. [DOI] [PubMed] [Google Scholar]
  198. Xiao ZC, Revest JM, Laeng P, Rougon G, Schachner M, Montag D. Defasciculation of neurites is mediated by tenascin-R. and its neuronal receptor F3/11. J. Neurosci. Res. 1998;52:390–404. doi: 10.1002/(SICI)1097-4547(19980515)52:4<390::AID-JNR3>3.0.CO;2-4. [DOI] [PubMed] [Google Scholar]
  199. Xu T, Bianco P, Fisher LW, Longenecker G, Smith E, Goldstein S, et al. Targeted disruption of the biglycan gene leads to an osteoporosis-like phenotype in mice. Nat. Genet. 1998;20:78–82. doi: 10.1038/1746. [DOI] [PubMed] [Google Scholar]
  200. Yamada H, Watanabe K, Shimonaka M, Yamaguchi Y. Molecular cloning of brevican, a novel brain proteoglycan of the aggrecan/versican family. J. Biol. Chem. 1994;269:10119–10126. [PubMed] [Google Scholar]
  201. Yamada H, Fredette B, Shitara K, Hagihara K, Miura R, Ranscht B, et al. The brain chondroitin sulfate proteoglycan brevican associates with astrocytes ensheathing cerebellar glomeruli and inhibits neurite outgrowth from granule neurons. J. Neurosci. 1997;17:7784–7795. doi: 10.1523/JNEUROSCI.17-20-07784.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  202. Yamagata M, Shinomura T, Kimata K. Tissue variation of two large chondroitin sulfate proteoglycans (PG-M/versican and PG-H/aggrecan) in chick embryos. Anat. Embryol. (Berlin) 1993;187:433–444. doi: 10.1007/BF00174419. [DOI] [PubMed] [Google Scholar]
  203. 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]
  204. Yick LW, Cheung PT, So KF, Wu W. Axonal regeneration of Clarke's neurons beyond the spinal cord injury scar after treatment with chondroitinase ABC. Exp. Neurol. 2003;182:160–168. doi: 10.1016/s0014-4886(02)00052-3. [DOI] [PubMed] [Google Scholar]
  205. Zaremba S, Guimaraes A, Kalb RG, Hockfield S. Characterization of an activity-dependent, neuronal surface proteoglycan identified with monoclonal antibody Cat-301. Neuron. 1989;2:1207–1219. doi: 10.1016/0896-6273(89)90305-x. [DOI] [PubMed] [Google Scholar]
  206. Zhang H, Kelly G, Zerillo C, Jaworski DM, Hockfield S. Expression of a cleaved brain-specific extracellular matrix protein mediates glioma cell invasion In vivo. J. Neurosci. 1998;18:2370–2376. doi: 10.1523/JNEUROSCI.18-07-02370.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Zhou XH, Brakebusch C, Matthies H, Oohashi T, Hirsch E, Moser M, et al. Neurocan is dispensable for brain development. Mol. Cell Biol. 2001;21:5970–5978. doi: 10.1128/MCB.21.17.5970-5978.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  208. Zuo J, Neubauer D, Graham J, Krekoski CA, Ferguson TA, Muir D. Regeneration of axons after nerve transection repair is enhanced by degradation of chondroitin sulfate proteoglycan. Exp. Neurol. 2002;176:221–228. doi: 10.1006/exnr.2002.7922. [DOI] [PubMed] [Google Scholar]

Articles from Journal of Anatomy are provided here courtesy of Anatomical Society of Great Britain and Ireland

RESOURCES