Skip to main content
Cold Spring Harbor Perspectives in Biology logoLink to Cold Spring Harbor Perspectives in Biology
. 2010 Jul;2(7):a001867. doi: 10.1101/cshperspect.a001867

Guidance Molecules in Axon Regeneration

Roman J Giger 1, Edmund R Hollis II 2, Mark H Tuszynski 3
PMCID: PMC2890195  PMID: 20519341

Abstract

The regenerative capacity of injured adult mammalian central nervous system (CNS) tissue is very limited. Disease or injury that causes destruction or damage to neuronal networks typically results in permanent neurological deficits. Injury to the spinal cord, for example, interrupts vital ascending and descending fiber tracts of spinally projecting neurons. Because neuronal structures located proximal or distal to the injury site remain largely intact, a major goal of spinal cord injury research is to develop strategies to reestablish innervation lost as a consequence of injury. The growth inhibitory nature of injured adult CNS tissue is a major barrier to regenerative axonal growth and sprouting. An increasing complexity of molecular players is being recognized. CNS inhibitors fall into three general classes: members of canonical axon guidance molecules (e.g., semaphorins, ephrins, netrins), prototypic myelin inhibitors (Nogo, MAG, and OMgp) and chondroitin sulfate proteoglycans (lecticans, NG2). On the other end of the spectrum are molecules that promote neuronal growth and sprouting. These include growth promoting extracellular matrix molecules, cell adhesion molecules, and neurotrophic factors. In addition to environmental (extrinsic) growth regulatory cues, cell intrinsic regulatory mechanisms exist that greatly influence injury-induced neuronal growth. Various degrees of growth and sprouting of injured CNS neurons have been achieved by lowering extrinsic inhibitory cues, increasing extrinsic growth promoting cues, or by activation of cell intrinsic growth programs. More recently, combination therapies that activate growth promoting programs and at the same time attenuate growth inhibitory pathways have met with some success. In experimental animal models of spinal cord injury (SCI), mono and combination therapies have been shown to promote neuronal growth and sprouting. Anatomical growth often correlates with improved behavioral outcomes. Challenges ahead include testing whether some of the most promising treatment strategies in animal models are also beneficial for human patients suffering from SCI.


Growth inhibition mechanisms active in the adult CNS limit our ability to repair spinal chord injuries. Promising new therapies aim to attenuate these mechanisms and activate growth-promoting programs.

THE REGENERATIVE CAPACITY OF INJURED CENTRAL NERVOUS SYSTEM IS LIMITED

In higher vertebrates, including humans, the regenerative capacity of neurons in the injured adult central nervous system (CNS) is extremely limited. Depending on the location and severity of the injury, trauma to the CNS may cause substantial damage to nervous system tissue that results in permanent neurological deficits. In the spinal cord, for example, injury often results in an interruption of vital ascending and descending pathways causing a range of functional deficits. The long-term goal of spinal cord injury (SCI) research is to develop strategies to improve or restore these deficits. One key step toward this goal is to reestablish neuronal innervation interrupted by SCI.

Reinnervation may be established by one of three strategies: (Fig. 1A) long-distance axonal regeneration followed by synapse formation on appropriate (pre-injury) target cells; (Fig. 1B) short-distance axonal regeneration and synapse formation to create relays to distal targets; or (Fig. 1C) sprouting of spared axons that maintain connectivity beyond the injury site (Fig. 1). Interestingly, evidence suggests that the limited spontaneous recovery that is observed following CNS injury is most likely a result of sprouting and compensation from spared systems. As discussed below, long-distance axon regeneration often occurs following peripheral nervous system (PNS) injury but does not occur spontaneously in the injured adult CNS. Thus, in mammals, injured neurons of the PNS and CNS show quite distinct adaptive strategies to injury. The disparity between neuronal responses following PNS and CNS injury is due in part to both intrinsic (cell-autonomous) and extrinsic factors.

Figure 1.

Figure 1.

Strategies to reestablish neuronal innervation following injury. (A) Long-distance axonal regeneration. Following axon transection, the distal segment of the nerve undergoes Wallerian degeneration (dotted black lines). New axons (red) sprout from the proximal axon segment (blue) and reestablish synaptic contact with preinjury targets. (B) short-distance growth of injured axons. Collateral sprouts (red) form synaptic contact with neighboring neuronal elements to by-pass the injury site. (C) Sprouting of spared axons to maintain connectivity beyond the injury site. The strategy shown in A is typically observed following compression injury in the PNS, whereas neuronal responses shown in B and C, have been observed in the injured adult mammalian CNS.

In this article we focus on the role of axon guidance molecules in the mature CNS with an emphasis on regeneration following injury. We also discuss two classes of inhibitory molecules, the chondroitin sulfate proteoglycans (CSPGs), and the prototypic myelin inhibitors MAG, Nogo, and OMgp. Both classes of molecules are profoundly inhibitory for neurite extension in vitro, and evidence suggests they also restrict neuronal growth and plasticity in the injured adult mammalian CNS in vivo. We discuss the role of neurotrophic factors and growth-permissive molecules in the adult CNS, and provide evidence for their growth-promoting effects in models of CNS regeneration.

WHY DON'T SEVERED CNS AXONS UNDERGO SPONTANEOUS REPAIR?

The regenerative capacity of the PNS and CNS is remarkably different. Following PNS injury, sensory and motor axons can and often do regenerate over long-distances, supporting substantial axonal regeneration and functional recovery. Why is regenerative axonal growth in the CNS so limited? Thirty years ago, an elegant series of transplantation experiments by Aguayo and colleagues established that some populations of adult CNS neurons possess the capacity to extend long axons following injury when provided with a favorable growth environment. In the presence of a peripheral nerve graft, CNS neurons could extend axons over several centimeters into the grafted tissue (Aguayo et al. 1981). We now know, however, that this capacity does not extend to the most important motor system for primate motor function, the corticospinal projection (Grill et al. 1997; Hollis et al. 2009a). Conversely, optic nerves transplanted into the PNS poorly support growth of denervated sciatic nerves. Few sciatic nerve axons entered the optic nerve transplant, whereas most of them bypassed the transplant before reentering the distal peripheral nerve stump (Aguayo et al. 1978). Collectively, nerve transplantation experiments uncovered two important principles for axonal regeneration: (1) some populations of CNS neurons retain a capability for long-distance axon growth throughout adulthood, and (2) the PNS milieu, but not the CNS milieu, is conducive for long-distance axon regeneration in vivo.

Subsequent studies revealed that CNS myelin formed by mature oligodendrocytes is profoundly inhibitory for neurite outgrowth. CNS myelin contains several factors that inhibit neurite outgrowth (Caroni and Schwab 1988), including the inhibitors myelin-associated glycoprotein (MAG), the reticulon family member RTN4a/Nogo-A, and oligodendrocyte myelin glycoprotein (OMgp) (Filbin 2003; Schwab 2004; Yiu and He 2006). In addition, several growth inhibitory molecules belonging to families of canonical axon guidance molecules are found in CNS myelin, including members of the netrin, ephrin, and semaphorin families (Bolsover et al. 2008; Löw et al. 2008). Although the growth inhibitory nature of CNS myelin is well established, regenerating CNS axons are faced with a number of additional obstacles. Within days following injury, a glial scar forms around the injury composed of reactive astrocytes, microglia, and meningeal fibroblasts that migrate into the lesion site. This “scar” formed at the lesion is thought to pose a physical barrier to axonal regeneration (Fig. 2). In addition, scar-associated molecules, including CSPGs, function as chemical inhibitors that block axon regeneration (Bradbury et al. 2002). Thus, the extensive expression of multiple classes of inhibitory molecules in injured CNS tissue is believed to constitute a major hurdle for regenerating axons (Table 1).

Figure 2.

Figure 2.

Schematic of adult rat spinal cord with dorsal lesion cavity as a result of injury. Transected axons of ascending sensory fibers (green) display dystrophic end-bulbs (inset) at the edge of the lesion cavity. The lesion cavity is surrounded by a glial scar (red) composed of reactive astrocytes, microglia and meningeal fibroblasts that migrate into the lesion site. Growth inhibitory molecules, including semaphorins and chondroitin sulfate proteoglycans (CSPGs) are enriched in the glial scar contributing to the growth inhibitory nature of injured adult mammalian CNS tissue.

Table 1.

comparison of factors that influence axonal regeneration following injury to the adult mammalian peripheral (PNS) and central (CNS) nervous system.

Nervous system injury PNS CNS
Spontaneous axonal regeneration Good: substantial regeneration often occurs following compressive injury Poor: very limited and mostly incomplete regeneration of severed fibers; typically results in permanent functional deficits distal to the injury
Environment Growth permissive Nonpermissive
-myelin Contains inhibitory molecules, however some of the known CNS myelin inhibitors are less abundant or not present in PNS myelin. Contains multiple inhibitory factors, including MAG, Nogo, OMgp, netrin-1, Sema4D, and ephrinB3 among others.
-Wallerian degeneration Occurs rapidly following injury, myelin debris is cleared from distal stump of nerve by Schwann cells and activated macrophages Occurs slowly, protracted and incomplete clearance of myelin debris from distal portion of injured fiber tracts
-glial scar No glial scar is formed at the injury site, however CSPGs are present Forms within days of injury and may inhibit axonal regeneration, rich in inhibitory CSPGs
-growth factors abundant, up-regulated following injury not expressed in temporal or spatial gradients supportive of regeneration
Immune system response Supports clearance of myelin debris in distal nerve stump Prolonged, with recruitment of innate and possibly adaptive immune cells
Cell intrinsic mechanisms
-rate of axonal regrowth Possess the ability to regenerate axons throughout adulthood at a rate of ∼1 mm per day. Some neurons possess the ability to regrow axons throughout adulthood if provided with a growth permissive environment.
- form of process growth Capability for long-distance axonal growth. Injury-induced neuronal plasticity results in reactive sprouting of processes from injured and noninjured neurons. Long-distance axonal regeneration does not occur.
Secondary damage Minimal Extensive degeneration after the injury that substantially contributes to parenchymal destruction
Cell-intrinsic growth programs Activated and efficient Deficient in various CNS populations, esp. in upper motoneurons that form the corticospinal tract
“Bridges” for regeneration in lesion site Spontaneously formed by Schwann cells, macrophages and fibroblasts Absent, resulting in failure of axonal attachment and extension

AXON GUIDANCE MOLECULES REGULATE NEURONAL STRUCTURE BEYOND THE INITIAL PHASE OF NEURONAL NETWORK ASSEMBLY

Following nervous system development, expression patterns of numerous axon guidance molecules are decreased, or altered, whereas others retain embryonic expression levels and are present in abundance in the mature brain and spinal cord. The expression of these molecules in the adult implies additional roles for guidance cues beyond the initial phase of neuronal process outgrowth, growth cone navigation, and target innervation. Recent evidence indicates that axon guidance molecules participate in a number of network refinement processes that occur after the initial scaffold of connectivity has been established (Bagri et al. 2003; Morita et al. 2006; Fu et al. 2007; Paradis et al. 2007; Low et al. 2008; Xu and Henkemeyer, 2009). In addition, guidance cues can regulate aspects of neuronal excitability and synaptic function in the mature CNS (Klein, 2009; Pasterkamp and Giger 2009). Once fully developed, neuronal circuits in the mature nervous system become more stable; however, it is also important to point out that adult neuronal connectivity is not hardwired. Indeed, more restricted forms of structural plasticity persist throughout adulthood in response to experience, injury, and aging. Because many mature neurons continue to express receptors for guidance cues, it has been speculated that inhibitory and chemorepulsive axon guidance molecules play a role in synaptic stabilization and limitation of neuronal plasticity in adulthood.

Of significance for studies on nervous system regeneration is the up-regulation of the expression of many guidance cues with known inhibitory activity, including members of the semaphorin, ephrin, netrin, Wnt, and slit families. Conversely, neurotrophic factors and permissive guidance cues are thought to promote neuronal growth and structural changes in adulthood (Sofroniew et al. 1990). A tightly regulated balance between growth-promoting and growth-inhibiting molecules is likely to determine the extent and type of neuronal structural changes that may occur in the mature CNS.

CANONICAL AXON GUIDANCE MOLECULES CONTRIBUTE TO THE GROWTH INHIBITORY NATURE OF INJURED ADULT MAMMALIAN CNS TISSUE

Although a great deal is known about extracellular molecules and signaling pathways that regulate axonal growth and pathfinding during development, comparatively little is known about the mechanisms that regulate neuronal growth and plasticity following injury to the adult nervous system. A number of axon guidance molecules are expressed in the adult CNS and their expression is regulated following injury. When coupled with the observation that many CNS neurons continue to express guidance receptors, this implies that adult CNS neurons remain responsive to guidance cues throughout life. Ironically, the picture emerging from these studies is that the inability of severed axons to undergo spontaneous repair in the adult CNS is, at least in part, attributable to the presence of the very same molecules that were so important during development in establishing the network. Below we summarize experimental evidence for a role of axon guidance molecules in CNS regeneration and provide specific examples for some of the best characterized guidance cues.

SEMAPHORINS

Many semaphorins function as inhibitory or repulsive guidance cues, and the presence of these proteins in the mature brain and spinal cord suggests roles in network stabilization by limiting neuronal growth. Indeed, similar to embryonic DRG neurons, adult DRG neurons show growth cone collapse in the presence of acutely applied Sema3A in vitro (Reza et al. 1999), and preconditioned adult DRG axons stop growing on encountering cells in SCI lesion sites that express Sema3A (Pasterkamp et al. 2001). Class 3 semaphorins (Sema3s) are expressed by glial scar-associated meningeal cells and have been proposed to contribute to the growth inhibitory nature of injured CNS tissue (Pasterkamp and Verhaagen, 2006). Interfering with the interaction between Sema3s and CSPGs blocks Sema3A repulsion in vitro, raising the possibility that Sema3s secreted by meningeal cells augment inhibition by glial scar tissue in a CSPG-dependent manner (Pasterkamp and Verhaagen 2006).

Recently, a small molecule agent (SM-216289) was found to block binding of Sema3A to the neuropilin-1/plexinA receptor complex, attenuating Sema3A repulsion of DRG neurons in vitro (Kikuchi et al. 2003). Further, SM-216289 accelerates axon regeneration in a rat model of olfactory nerve axotomy (Kikuchi et al. 2003), and it has been reported to enhance growth of neuropilin-1-expressing serotonergic axons after SCI in rats (Kaneko et al. 2006). In the same injury model, blocking Sema3A signaling does not lead to enhanced regeneration of corticospinal axons or ascending sensory axons (Kaneko et al. 2006), suggesting that blocking Sema3A function enhances growth of a subset of axons. In organotypic brain slices, transection of the entorhinal-hippocampal pathway (EHP) leads to up-regulation of Sema3A and neuropilin-1 expression in the hippocampus and entorhinal cortex. No spontaneous regeneration of severed EHP axons is observed. In the presence of a peptoid inhibitor that selectively blocks the Sema3A-neuropilin-1 interaction, the number of EHP axons that grows into the denervated hippocampus increases significantly (Montolio et al. 2009). Together these studies support the idea that Sema3A inhibits regenerative axonal growth in vitro and in vivo. As Sema3A (and other class 3 semas) not only regulate neuronal growth but also play important roles in vascular remodeling (Wang et al. 2005), immune system function (Suzuki et al. 2008) (Mizui et al. 2009), and cell death (Bagnard et al. 2004; Giraudon et al. 2004; Ben-Zvi et al. 2008; Moretti et al. 2008), any of these activities could influence outcomes after SCI. Additional studies are needed to more precisely define the mechanism(s) and role of Sema3s in the injured CNS.

Growing evidence suggests that in addition to Sema3s, membrane-associated semaphorins contribute to the regenerative failure of injured CNS axons. Sema4D, expressed by oligodendrocytes and transiently up-regulated near sites of spinal cord injury, inhibits outgrowth of postnatal cerebellar and sensory neurites in vitro (Moreau-Fauvarque et al. 2003). Similarly, Sema7A is expressed by oligodendrocytes in spinal cord white matter (Pasterkamp et al. 2007) and Sema6B is strongly up-regulated near the lesion site following transection of the fornix in the adult rat (Kury et al. 2004). Whether targeting of membrane-bound semaphorins will influence outcomes following CNS injury is an important question for future studies.

EPHRINS

Similar to the semaphorins, the predominant neuronal response to ephrins is repulsive. Ephrins bind to members of the EphA and EphB receptor tyrosine kinase families. Expression of several ephrins and Eph receptors continues beyond nervous system development and remains robust in the mature rodent (Liebl et al. 2003) and human (Sobel, 2005) CNS. Of interest for nervous system regeneration is the strong expression of ephrinB3 in CNS myelin, the injury-induced up-regulation of ephrinB2 in reactive astrocytes, and the increase in ephrinA5 expression around ischemic cortical lesions (Bundesen et al. 2003; Benson et al. 2005; Carmichael et al. 2005). In vitro, ephrinB3 is a strong inhibitor of neurite outgrowth for postnatal cortical neurons and functions in a EphA4-dependent manner (Benson et al. 2005). In spinal cord injured rats, EphA4 protein accumulates in severed CST axons, suggesting that they are responsive to ephrin ligands present in myelin (ephrinB3) and scar tissue (ephrinB2) (Fabes et al. 2006). It was found that blocking of EphA4 with an infused peptide agonist enhances sprouting of CST axons rostral to the injury site but fails to promote axonal regeneration across the lesion into the distal portion of the spinal cord (Fabes et al. 2007). Although a regeneration phenotype was reported for spinal cord injured EphA4 null mice through a spinal cord hemisection lesion site (Goldshmit et al. 2004), lesion completeness could not be determined with confidence in this report. The exact mechanism by which loss of EphA4 may influence axonal growth in the adult CNS is complicated by the observation that the protein serves ligand as well as receptor functions in neurons and glia. EphA4 expressed by CST axons may function as an inhibitory receptor for ephrinB3 and ephrinB2. Moreover, EphA4 is up-regulated by reactive astrocytes and strongly inhibits neurite outgrowth through a reverse signaling mechanism (Goldshmit et al. 2004). In EphA4 mutant mice, reactive gliosis and expression of CSPGs is reportedly reduced compared with wild-type mice (Goldshmit et al. 2004). Additional evidence that ephrin-Eph signaling may play a role in CNS response to injury stems from reports showing up-regulation of EphA3 in astrocytes after SCI in rats (Irizarry-Ramirez et al. 2005). Similarly, EphA7 is up-regulated by SCI and is thought to be a regulator of apoptosis in rat astrocytes (Figueroa et al. 2006). Thus, in addition to their role as growth inhibitory cues, Eph-ephrin signaling also influences formation of the glial scar and apoptosis.

WNTS

Another set of developmental guidance molecules implicated in SCI is the Wnt family. Decreasing anterior to posterior gradients of Wnts mediate both anterior growth of postcrossing commissural axons as well as posterior growth of descending corticospinal axons (Lyuksyutova et al. 2003; Liu et al. 2005). The differential response of ascending and descending axons to Wnt stimulation is thought to be mediated by differential receptor expression, with Frizzled receptors mediating Wnt-4 attraction and the atypical receptor tyrosine kinase Ryk promoting Wnt-5a-mediated posterior growth of the corticospinal tract through repulsion (Lyuksyutova et al. 2003; Liu et al. 2005). Little is known of the signaling cascades downstream of the Wnt-Ryk interaction, although there is evidence to suggest that Ryk and Frizzled act as function-modulating Wnt coreceptors (Lu et al. 2004b; Li et al. 2009). Following dorsal column injury, re-induction of Wnt-5a surrounding the lesion correlates with Ryk induction in corticospinal axons and axonal die-back (Liu et al. 2008; Miyashita et al. 2009). Infusion of functional blocking Ryk antibodies appears to either reduce axonal die-back or promote sprouting of lesioned corticospinal axons (Liu et al. 2008). Similar to re-induction of Ryk in injured corticospinal neurons, Ryk expression is up-regulated in DRG neurons following peripheral nerve injury (Song et al. 2008), although DRG neurons nonetheless show an extensive capacity for regeneration.

PROTOTYPIC MYELIN INHIBITORS: NOGO, MAG, AND OMgp

Growth inhibitory molecules that do not belong to any of the known families of axon guidance molecules have been identified. These include the myelin-associated inhibitors Nogo-A, MAG, and OMgp, hereafter called the prototypic myelin inhibitors (Fig. 3). Nogo-A (RTN4a) is a member of the reticulon (RTN) family of membrane associated proteins (Chen et al. 2000; GrandPre et al. 2000; Prinjha et al. 2000) and is comprised of two distinct growth inhibitory domains: 1) Amino-Nogo, an activity that inhibits both neurite outgrowth and the adhesion of several nonneuronal cell types, and 2) Nogo66, a 66-amino acid residue hydrophilic loop (Fig. 3) (Fournier et al. 2001; Oertle et al. 2003). Antibody blocking of Nogo-A in rats has been reported to facilitate long-distance regeneration and sprouting of corticospinal axons (Schnell and Schwab, 1990; Bregman et al. 1995; Thallmair et al. 1998). Growth of serotonergic fibers in the presence of anti-Nogo-A has also been reported (Gonzenbach and Schwab, 2008). However, regeneration studies in mice null for Nogo did not show enhanced longitudinal growth of severed CST axons (Zheng et al. 2003; Lee et al. 2009b); similar to wild-type mice, CST axons of Nogo null mice fail to extend past the injury site. Together with other studies (GrandPre et al. 2002; Kim et al. 2003; Simonen et al. 2003; Dimou et al. 2006; Steward et al. 2008), the overall effectiveness of Nogo-neutralizing approaches for SCI remains the subject of some debate: It is clear that Nogo inhibits axonal growth, but the ability of Nogo neutralization alone to facilitate axonal sprouting or regeneration, in the presence of a number of other myelin and ECM-associated inhibitors, remains uncertain (Zheng et al. 2005; Lee et al. 2009b). Nonetheless, Nogo neutralizing antibody infusions are now undergoing translational human testing in acute SCI.

Figure 3.

Figure 3.

The prototypic myelin inhibitors MAG, Nogo-A, and OMgp and their receptors. MAG (Siglec 4a) is a type-1 transmembrane protein with an ectodomain composed of five Ig-like domains. Nogo-A (RTN4A) has two inhibitory domains: Amino-Nogo and Nogo-66. The membrane topology of Nogo-A appears to be variable. Although Amino-Nogo can be detected extracellulary, a significant portion of Amino-Nogo is thought to have a cytoplasmic orientation (dotted line). OMgp is a member of the large family of leucine-rich repeat (LRR) proteins and linked to the cell surface by a GPI anchor.

Several receptors for prototypic myelin inhibitors have been identified: PirB (and its human homolog LILRB2) binds to MAG, Nogo-66, and OMgp and signals neuronal growth cone collapse and neurite outgrowth inhibition in vitro. The ectodomain of PirB is composed of Ig-like domains and the cytoplamsic portion contains four immunoreceptor tyrosine-based inhibitory motifs (ITIM). The Nogo-66 receptor 1 (NgR1) binds directly to Nogo-66, MAG, and OMgp and is important for growth cone collapse responses toward acutely presented inhibitors. In some neurons, NgR1 is thought to form a tripartite receptor complex with Lingo-1 and the death domain containing TNFR superfamily members p75 or its functional substitute TROY (TNFRSF19). NgR2 is a receptor selective for MAG, functionally redundant with NgR1. The lectin-activity of MAG forms a complex with select gangliosides, including GT1b. Integrins, a family of heterodimeric cell surface receptors composed of α- and β-subunits, have been found to participate in Nogo-A and MAG inhibition in vitro.

Myelin-associated glycoprotein (MAG) is a sialic acid-binding Ig-superfamily lectin (siglec4a) composed of 5 Ig-like domains, a single transmembrane domain and a short cytoplasmic domain (Fig. 3) (Filbin, 2003). In vitro, MAG regulates neurite outgrowth in an age-dependent manner. MAG promotes growth of many types of embryonic and neonatal neurons (Johnson et al. 1989) and, at more mature stages, inhibits neurite outgrowth from a broad spectrum of primary neurons (DeBellard et al. 1996). The MAG lectin activity, located within the first two Ig-like domains, binds to a broad range of sialoglycans, including ganglioside GT1b, and has been found to augment the neurite outgrowth inhibitory activity of soluble MAG in some neuronal populations in vitro (Vinson et al. 2001; Vyas et al. 2002). When presented in membrane bound form, the MAG lectin activity is largely dispensable for neurite outgrowth inhibition (Tang et al. 1997) and can be dissociated from the MAG growth inhibitory site (Cao et al. 2007; Robak et al. 2009; Worter et al. 2009). Although the growth inhibitory nature of MAG is well established, mice carrying a null allele for MAG do not show enhanced growth of injured corticospinal or optic nerve axons when compared with wild-type controls (Bartsch et al. 1995).

The third molecule of the prototypic myelin inhibitors is OMgp, a member of the leucine-rich repeat (LRR) protein family (Wang et al. 2002b). OMgp is linked via a glycosylphosphatidylinositol (GPI) anchor to the cell surface and is expressed by myelinating glia in the CNS but not the PNS (Mikol and Stefansson, 1988). In addition, OMgp is strongly expressed by many types of neurons in the mature CNS (Habib et al. 1998; Lee et al. 2009a). Although OMgp null mice do not show detectably enhanced growth of axotomized corticospinal axons, increased sprouting of serotonergic axons has been reported (Ji et al. 2008).

Taken together, deletion or neutralization of Nogo, MAG, or OMgp alone results in limited or no regeneration of corticospinal axons, the most important system for voluntary motor control in humans. Other axonal systems, including descending serotonergic or raphespinal axons, may show apparent genetic-background dependent increases in axonal growth in mutants compared with wild-type controls.

MECHANISMS OF MYELIN-ASSOCIATED GROWTH INHIBITION

The first mechanistic clue regarding the function of the prototypic myelin inhibitors stemmed from the identification of the Nogo66 receptor 1 (NgR1) as a high affinity receptor for the Nogo inhibitory peptide Nogo66, MAG and OMgp (Fig. 3) (Fournier et al. 2001; Domeniconi et al. 2002; Liu et al. 2002a; Wang et al. 2002b). The NgR1 related molecule NgR2, supports binding of MAG, but unlike NgR1, does not associate with Nogo66 or OMgp (Venkatesh et al. 2005). Functional studies with primary neurons obtained from NgR1 null mice revealed that NgR1 is necessary for collapse of growth cones in postnatal dorsal root ganglion neurons following acute presentation of soluble Nogo66, MAG, or OMgp (Kim et al. 2004; Chivatakarn et al. 2007). When plated on substrate bound Nogo66, MAG, or OMgp, however, various types of primary neurons null for NgR1 are strongly inhibited, and growth inhibition is comparable to wild-type neurons from littermate controls (Zheng et al. 2005; Chivatakarn et al. 2007; Venkatesh et al. 2007). The combined loss of NgR1 and NgR2 leads to a partial disinhibition of DRG neurons cultured on fibroblasts stably expressing MAG (Worter et al. 2009). NgR1 has been reported to complex with Lingo-1 and select members of the tumor necrosis factor receptor (TNFR) superfamily to signal growth inhibition in vitro (Fig. 3) (Wang et al. 2002a; Mi et al. 2004; Park et al. 2005; Shao et al. 2005). Growth inhibitory responses to Amino-Nogo are not well understood; indirect and integrin-dependent mechanisms for Amino-Nogo mediated inhibition have been reported (Hu and Strittmatter, 2008).

More recently, paired immunoglobulin-like receptor B (PirB), and its human homolog (LILRB2), were identified as novel receptors for Nogo-66, OMgp, and MAG (Atwal et al. 2008). PirB, a member of the leukocyte immunoglobulin receptor (LIR) subfamily, is comprised of six Ig-like domains, a transmembrane segment, and a cytoplasmic region harboring immunoreceptor tyrosine-based inhibitory motifs (Fig. 3). Antibody blocking or genetic ablation of PirB renders primary neurons more resistant to inhibition by substrate bound Nogo66, OMgp, MAG, or CNS myelin. The combined blockade of NgR1 and PirB largely abolishes neurite outgrowth inhibition on substrate bound CNS myelin (Atwal et al. 2008). These experiments reveal a significant degree of functional redundancy for the mechanisms used by prototypic myelin inhibitors. Furthermore, PirB and NgR1 are functionally linked and collaborate in signaling neurite outgrowth inhibition (Atwal et al. 2008).

Consistent with the idea that there is significant functional redundancy among the receptor systems used by myelin inhibitors, CST axons in spinal cord injured NgR1 null mice fail to grow past the lesion site (Kim et al. 2004; Zheng et al. 2005). Because of its direct interaction with MAG, Nogo-66 and OMgp, a soluble peptide of the NgR1 ligand binding domain was developed (NgR1(310)-Fc) and used to antagonize myelin inhibition. NgR1(310)-Fc complexes with NgR1 ligands and competes for ligand binding to neuronal cell surface receptors, including NgR1. In vitro, NgR1(310)-Fc overcomes CNS myelin inhibition and promotes neurite outgrowth of different types of neurons plated on substrate bound CNS myelin or individual inhibitors (Fournier et al. 2002; He et al. 2003; Zheng et al. 2005; Peng et al. 2009). Interpretation of in vivo effects of NgR1(310)-Fc administration is more complicated: intrathecally administrated NgR1(310)-Fc was reported to increase growth of corticospinal and raphaespinal axons and to improve functional outcome after spinal cord injury (Li et al. 2004; Wang et al. 2006). However, at least some of the reported effects of NgR1(310)-Fc may be an artifact of axon labeling methods (Steward et al. 2007). In an independent study, NgR1(310)-Fc was reported to stimulate regrowth of myelinated sensory axons into the dorsal root entry zone, spinal cord white and grey matter following dorsal root crush injury (Harvey et al. 2009; Peng et al. 2009).

EXPERIENCE-DEPENDENT AND INJURY-INDUCED NEURONAL PLASTICITY ARE REGULATED BY RELATED MECHANISMS

In the immune system, PirB and its close relative PirA are receptors for major histocompatibility complex (MHC) class I molecule(s) (Takai 2005). MHC class I proteins are ubiquitously expressed and include classical and nonclassical molecules essential for adaptive and innate immune responses. In the developing and mature CNS select MHC class I molecules show distinct neuronal distribution patterns, and expression is regulated by neuronal activity (Corriveau et al. 1998). In the hippocampus and visual system, MHC class I molecules regulate activity-dependent changes in synaptic connectivity (Huh et al. 2000). Perturbations of MHC class I function in the hippocampus enhance long-term potentiation (LTP) at Schaffer collateral-CA1 synapses (Huh et al. 2000). Similar to MHC class 1 molecules, PirB is expressed in CNS neurons and participates in limiting neuronal plasticity. In mice lacking transmembrane anchored PirB (called PirBTM mice), cortical ocular dominance (OD) plasticity is more robust at all ages compared with wild-type controls (Syken et al. 2006). Interestingly, defects observed in the visual system of PirBTM, NgR1, and Nogo-A/B mutant mice are very similar: connectivity in the visual cortex of mutant mice is not consolidated at the end of the critical period and OD plasticity is more robust in adulthood (McGee et al. 2005; Syken et al. 2006). In the mature hippocampus, NgR1 is found at synapses and influences dendritic spine morphology in vivo (Lee et al. 2008). NgR1 not only influences neuronal structure but also modulates synaptic strength at Schaffer collateral-CA1 synapses (Lee et al. 2008). Interestingly, some of the synaptic defects reported for NgR1 mutants resemble those reported in mice deficient for MHC class I molecules (Huh et al. 2000). Together, these findings reveal insights into the physiological function of myelin inhibitors and also suggest that mechanisms influencing neuronal growth following CNS injury and also synaptic plasticity in the intact CNS are related.

Because of the growing complexity of molecular players contributing to the growth inhibitory milieu of injured CNS tissue, plasticity from spared fiber tracts is a potentially more tractable target for improving functional outcome after SCI than true axonal regeneration. Indeed, SCI results in collateral sprouting of corticospinal axons with formation of novel connections to intraspinal neurons (Bareyre et al. 2004). Similarly, injury-induced sprouting of spared reticulospinal axons, another modulator of motor function, has been described (Ballermann and Fouad 2006). Most recently, extensive synaptic rearrangements of propriospinal projections after spinal cord injury, leading to improvement in locomotion, has been reported (Courtine et al. 2008). When coupled with the fact that many humans with complete loss of function below a site of SCI nonetheless show spared axons in rims of peripheral white matter (Tuszynski et al. 1999), enhancement of plasticity from spared axons is an alternative and compelling target for SCI therapy.

EXTRACELLULAR MATRIX MOLECULES AND AXONAL GROWTH AFTER INJURY

An important class of inhibitory ECM molecules are the chondroitin sulfate proteoglycans (CSPGs), a diverse group of glycoproteins composed of a core protein covalently linked to specific types of glycosaminoglyan (GAG) side chains (Galtrey and Fawcett, 2007). Several types of CSPGs are found in adult CNS tissue, many of which are expressed throughout the brain and spinal cord and condensed into perineuronal nets surrounding the somata and dendrites of various types of CNS neurons (Bruckner et al. 2000). CSPGs are up-regulated following CNS injury (McKeon et al. 1995; Davies et al. 1996; Fitch et al. 1999; Jones et al. 2002; Morgenstern et al. 2002; Jones et al. 2003b) and appear to inhibit neurite outgrowth from adult neurons (Snow et al. 1990; Davies et al. 1999). Enzymatic degradation of the GAG side chains of CSPGs, using chondroitinase ABC (ChABC), largely abrogates the neurite outgrowth inhibitory action of substrate bound CSPGs in vitro (McKeon et al. 1995; Zuo et al. 1998; Grimpe et al. 2005). In the rat visual cortex, local delivery of ChABC allows experience-dependent neuronal plasticity beyond the critical period (Pizzorusso et al. 2002) and reportedly enhances axonal growth following CNS injury (Moon et al. 2001; Bradbury et al. 2002; Tester and Howland 2008). More recent work shows that ChABC treatment in rats after SCI opens a window during which rehabilitative training supports functional improvement (Garcia-Alias et al. 2009). Importantly, only the trained skills are improved in injured animals, suggesting that during the window of enhanced plasticity the formation of new and appropriate connections may be driven by task-specific training. Rehabilitation only enhances functions that are trained and may come at a high cost, as other tasks that are not trained are worsened compared with animals that receive no training at all (Garcia-Alias et al. 2009). ChABC treatment reportedly also has neuroprotective effects: following SCI, administration of ChABC near the injury site prevents atrophy of axotomized corticospinal projection neurons following dorsal column lesion and activates growth promoting intracellular signaling pathways (Carter et al. 2008)

Members of the receptor protein tyrosine phosphatases (RPTPs) of the leukocyte antigen-related (LAR) subfamily have previously been shown to associate with heparan sulfate proteoglycans (HSPGs) (Aricescu et al. 2002; Fox and Zinn, 2005; Johnson et al. 2006). The family member RPTPσ was recently identified as a receptor that directly interacts with growth inhibitory CSPGs and signals neuronal inhibition toward neurocan, aggrecan, or CSPGs expressed on the surface of astrocytes in vitro (Fig. 4) (Shen et al. 2009). In the mature mouse PNS, loss of RPTPσ enhances regenerative axonal growth following injury to the sciatic or facial nerve (McLean et al. 2002; Thompson et al. 2003). In the CNS, loss of RPTPσ has been reported to promote axon regeneration in the injured optic nerve (Sapieha et al. 2005) and to markedly improve axon extension into the lesion penumbra following dorsal column crush injury (Shen et al. 2009). Effects of loss of RPTPσ on corticospinal axon regeneration reported in a recent study seem to be very robust but are more difficult to interpret, as lesions may have been incomplete (Fry et al. 2009).

Figure 4.

Figure 4.

Mechanisms for CSPG inhibition. Proteoglycans are a heterogenous class of extracellular proteins with distinct protein core structures bearing covalently attached sulfated glycosaminoglycan (GAG) side chains. Prominent members of neural proteoglycans bearing heparan sulfate GAG chains (HSPGs) include the syndecans, glycpicans and agrin. Chondroitin sulfate GAG bearing proteoglycans (CSPGs) include members of the lectican family (neurocan, aggrecan, versican, and brevican). The leukocyte antigen-related (LAR) subfamily of receptor protein tyrosine phosphatases (RPTPs) includes the mammalian members LAR, RPTPσ and RPTPδ. LAR type RPTPs are transmembrane proteins with cell adhesion molecule-like ectodomains and a large cytoplasmic region with two conserved phosphatase domains. The Drosophila homolog of LAR-RPTPσ-RPTPδ is called DLAR and is a functional receptor for the fly HSPGs syndecan and glypican (dally-like). Avian RPTPσ binds directly to the HSPGs agrin and collagen XVIII. In addition, CSPGs belonging to the lectican family (including neurocan and aggrecan) bind to the first Ig-like domain of mouse RPTPσ to signal neuronal growth inhibition. The interaction between lecticans and RPTPσ is direct, depends on the presence of CS-GAG chains, and is sensitive to chondroitinase ABC treatment.

Counter-intuitively, axons regenerating into cellular substrates placed in a spinal cord lesion site have also been observed to specifically associate with cells expressing CSPGs (Jones et al. 2003a; Lu et al. 2007). These CSPG-expressing cells in the lesion site are host Schwann cells that have migrated into the site of injury (Jones et al. 2003a). Notably, these Schwann cells simultaneously express NCAM, L1, and possibly other classically “permissive” ECM and cell adhesion molecules (Jones et al. 2003a). These findings reflect the complexity of the injured in vivo CNS environment: most likely, the success or failure of axonal regeneration in vivo represents the summation of various inhibitory and permissive factors. If the amount of inhibition present in the extracellular matrix and on myelin exceeds the stimulation derived from growth factors and permissive substrates in the environment, then growth will fail, as observed in the mature CNS. On the other hand, regeneration may succeed if the balance tips in favor of growth, as observed in the injured peripheral nerve. The activation and persistence of an active growth state in the injured neuron further contributes to the success or failure of adult axonal regeneration.

GROWTH FACTORS AS GROWTH AND GUIDANCE SIGNALS IN THE INJURED CNS

Neurotrophic factors contribute to growth, guidance and survival of several neuronal populations during development. Neurotrophins, including nerve growth factor (NGF), brain-derived neurotrophic factor (BDNF), neurotrophin-3 (NT-3), and neurotrophin-4 (NT-4) have unique functions in the CNS and are differentially regulated during developmental CNS maturation. Neurotrophin expression is regulated both spatially and temporally, generally decreasing as development proceeds. However, neurotrophin expression persists widely in adulthood in many CNS regions associated with functional plasticity (Maisonpierre et al. 1990).

Notably, successful axonal regeneration in the adult PNS is accompanied by rapid production of several growth factors by Schwann cells, including NGF, BDNF, IGF, CNTF, GDNF and others (Meyer et al. 1992; Sendtner et al. 1992; Curtis et al. 1993; Funakoshi et al. 1993; Glazner et al. 1994; Naveilhan et al. 1997; Shen et al. 1999; Höke et al. 2000; Costigan et al. 2002). Depletion of Schwann cell-derived BDNF reduces regeneration and remyelination of both sensory and motor neurons (Zhang et al. 2000; Song et al. 2008; Geremia et al. 2009). Motor neuron death following root avulsion in the adult is also ameliorated by growth factors (Novikov et al. 1995; Kishino et al. 1997; Novikov et al. 1997).

Growth factors elicit extensive growth of several axonal populations after SCI. BDNF elicits growth of lesioned raphaespinal, cerulospinal, rubrospinal, and reticulospinal motor axons into permissive growth matrices placed in sites of SCI (Liu et al. 1999; Jin et al. 2002; Lu et al. 2005). NT-3 promotes regeneration of lesioned dorsal column proprioceptive sensory axons (Zhang et al. 1998; Bradbury et al. 1999; Oudega and Hagg, 1999; Ramer et al. 2000; Lu et al. 2004a; Taylor et al. 2006; Alto et al. 2009). NGF promotes growth of nociceptive axons (Tuszynski et al. 1996; Grill et al. 1997) and may contribute to the spontaneous development of dysfunctional pain after SCI.

Notably, the corticospinal projection appears to be among the most refractory axonal systems from which to elicit experimental regeneration after SCI (Blesch and Tuszynski, 2009). For example, whereas insulin-like growth factor-I (IGF-I) promotes regeneration of coerulospinal and raphaespinal axons, and prevents axotomy-induced death of cortiocspinal motor neurons, it does not promote regeneration of corticospinal axons into a lesion cavity filled with a substrate that supports growth of other axonal systems (Hollis II et al. 2009a). Similarly, BDNF prevents axotomy-induced atrophy of rubrospinal neurons and promotes their regeneration into sites of SCI (Kobayashi et al. 1997; Jin et al. 2002; Kwon et al. 2002; Liu et al. 2002b; Kwon et al. 2007), and BDNF prevents corticospinal neuronal death after axotomy (Giehl and Tetzlaff 1996; Giehl et al. 2001; Lu et al. 2001); yet BDNF does not promote corticospinal axonal regeneration into a spinal cord or cortical lesion site (Lu et al. 2001). If, however, corticospinal neurons are genetically modified to overexpress the BDNF receptor trkB at the same time that BDNF is expressed in a cortical lesion site, corticospinal axons can be induced to regenerate (Fig. 5) (Hollis II et al. 2009b). These findings indicate that different populations of axons in the CNS show distinct patterns of growth factor sensitivity. Moreover, patterns of growth factor sensitivity can be modified by altering the neuron's repertoire of growth factor receptor expression.

Figure 5.

Figure 5.

Corticospinal regeneration induced by over-expression of the high-affinity BDNF receptor trkB. (A) Corticospinal motor neurons are retrogradely infected with adeno-associated virus expressing GFP (green), whereas lentivirus encoding trkB (red) is delivered directly to the motor cortex. Some cells will coexpress both GFP and trkB (yellow). (B) Traced GFP-immunoreactive corticospinal axons demonstrate regeneration into a subcortical lesion grafted with BDNF-secreting syngeneic fibroblast substrate (red outlines). (C) Regenerated corticospinal axon within the subcortical graft. (D–F) Other examples of regenerated corticospinal axons within subcortical BDNF-secreting grats after trkB over-expression demonstrate growth cone-like morphology (D, E) and associations with the vasculature; axons indicated with arrowheads in (F) wrap around a blood vessel (asterisk). Scale bars, 500um (A), 25um (C–F). Adapted from Hollis et al. PNAS (2009).

Recent findings indicate that axons can be induced to regenerate not just into, but beyond, sites of SCI when multiple mechanisms influencing axonal growth are experimentally manipulated (Lu et al. 2004a; Pearse et al. 2004; Fouad et al. 2005; Houle et al. 2006; Lu and Tuszynski 2008; Alto et al. 2009; Kadoya et al. 2009). In these studies, the environment of the lesioned CNS is rendered more permissive to growth by cell grafting or peripheral nerve grafting into the central lesion site. In addition, the endogenous growth state of the neuron in several studies must be activated using either a conditioning lesion or cAMP administration to the injured neuronal soma (Neumann et al. 2002; Qiu et al. 2002; Lu et al. 2004a; Pearse et al. 2004; Alto et al. 2009; Kadoya et al. 2009). Importantly, additional chemoattractive growth signals must be provided beyond the lesion site (Lu et al. 2004a; Alto et al. 2009; Kadoya et al. 2009). Under the latter circumstances, axons will regenerate into host spinal cord beyond the lesion site (Lu et al. 2004a; Alto et al. 2009). The chemotropic gradient can, over distances of several millimeters, guide regenerating axons to appropriate preinjury targets. Once in the target, regenerating axons form phenotypically correct synapses at the ultrastructural level, complete with presynaptic elements containing clusters of synaptic vesicles that are indistinguishable from the prelesioned state. Yet, the reconstituted neural circuitry is not electrophysiologically active, likely because of the absence of remyelination following injury (Alto et al. 2009). These findings remind us that the reconstitution of functional activity in mature neural systems after injury is a highly challenging endeavor. Developmental patterning of the nervous system occurs as a result of a delicately orchestrated series of genetic and epigenetic events coordinated between neurons and their environment, and over very short distances. Recapitulating this series of events in the large, injured adult CNS is, to say the least, challenging.

SUMMARY

Depending on severity and location, injury to the adult spinal cord causes substantial damage to neural tissue and is typically associated with permanent functional deficits. Over the past two decades enormous progress has been made in our understanding of the molecular and cellular events triggered by injury, providing insights into key mechanisms that contribute to tissue damage and regenerative failure of injured CNS neurons. One major goal of SCI research is to reestablish neuronal connectivity lost as a consequence of injury. In patients with incomplete SCI, reinnervation may be established by short distance axonal sprouting and formation of new synaptic contacts with neurons that bypass the injury site. When coupled with sprouting of spared axons beyond the injury site, this may allow reinnervation of preinjury targets. Neuronal growth and axonal sprouting of injured and noninjured CNS neurons may be achieved by lowering environmental growth inhibitory signals, enhancing growth promoting signals, activation of intrinsic growth programs, or some combination thereof. It is likely that treatments that enhance neuronal growth and plasticity need to be combined with task-specific rehabilitative training to strengthen and consolidate functionally meaningful connections in an activity-dependent manner. Indeed, experimentally enhanced neuronal plasticity combined with task-specific training regimes following SCI are reminiscent of activity-dependent refinement processes that occur in the visual system during development. During a temporally restricted window, enhanced plasticity combined with activity shapes neuronal structure. At the end of this critical period, appropriate connections are established and stabilized. The similarity between restricted neuronal growth following CNS injury and experience-dependent plasticity at the end of the critical period is further underscored by a striking convergence of molecular mechanisms that limit neuronal plasticity in both processes. These recent advances in the field of SCI research are based on rodent SCI models. Critical next steps are likely to include SCI experiments in larger animal models to further develop treatment strategies and if successful, develop protocols for human clinical trials.

Footnotes

Editors: Marc Tessier-Lavigne and Alex L. Kolodkin

Additional Perspectives on Neuronal Guidance available at www.cshperspectives.org

REFERENCES

  1. Aguayo AJ, David S, Bray GM 1981. Influences of the glial environment on the elongation of axons after injury: Transplantation studies in adult rodents. J Exp Biol 95: 231–240 [DOI] [PubMed] [Google Scholar]
  2. Aguayo AJ, Dickson R, Trecarten J, Attiwell M, Bray GM, Richardson P 1978. Ensheathment and myelination of regenerating PNS fibres by transplanted optic nerve glia. Neurosci Lett 9: 97–104 [DOI] [PubMed] [Google Scholar]
  3. Alto LT, Havton LA, Conner J, Hollis ER II, Blesch A, Tuszynski MH 2009. Chemotropic guidance facilitates axonal regeneration into brainstem targets and synapse formation after spinal cord injury. Nat Neurosci 12: 1106–1113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Aricescu AR, McKinnell IW, Halfter W, Stoker AW 2002. Heparan sulfate proteoglycans are ligands for receptor protein tyrosine phosphatase sigma. Mol Cell Biol 22: 1881–1892 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Atwal JK, Pinkston-Gosse J, Syken J, Stawicki S, Wu Y, Shatz C, Tessier-Lavigne M 2008. PirB is a functional receptor for myelin inhibitors of axonal regeneration. Science 322: 967–970 [DOI] [PubMed] [Google Scholar]
  6. Bagnard D, Sainturet N, Meyronet D, Perraut M, Miehe M, Roussel G, Aunis D, Belin MF, Thomasset N 2004. Differential MAP kinases activation during semaphorin3A-induced repulsion or apoptosis of neural progenitor cells. Mol Cell Neurosci 25: 722–731 [DOI] [PubMed] [Google Scholar]
  7. Bagri A, Cheng HJ, Yaron A, Pleasure SJ, Tessier-Lavigne M 2003. Stereotyped pruning of long hippocampal axon branches triggered by retraction inducers of the semaphorin family. Cell 113: 285–299 [DOI] [PubMed] [Google Scholar]
  8. Ballermann M, Fouad K 2006. Spontaneous locomotor recovery in spinal cord injured rats is accompanied by anatomical plasticity of reticulospinal fibers. Eur J Neurosci 23: 1988–1996 [DOI] [PubMed] [Google Scholar]
  9. Bareyre FM, Kerschensteiner M, Raineteau O, Mettenleiter TC, Weinmann O, Schwab ME 2004. The injured spinal cord spontaneously forms a new intraspinal circuit in adult rats. Nat Neurosci 7: 269–277 [DOI] [PubMed] [Google Scholar]
  10. Bartsch U, Bandtlow CE, Schnell L, Bartsch S, Spillmann AA, Rubin BP, Hillenbrand R, Montag D, Schwab ME, Schachner M 1995. Lack of evidence that myelin-associated glycoprotein is a major inhibitor of axonal regeneration in the CNS. Neuron 15: 1375–1381 [DOI] [PubMed] [Google Scholar]
  11. Ben-Zvi A, Manor O, Schachner M, Yaron A, Tessier-Lavigne M, Behar O 2008. The Semaphorin receptor PlexinA3 mediates neuronal apoptosis during dorsal root ganglia development. J Neurosci 28: 12427–12432 [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Benson MD, Romero MI, Lush ME, Lu QR, Henkemeyer M, Parada LF 2005. Ephrin-B3 is a myelin-based inhibitor of neurite outgrowth. Proc Natl Acad Sci 102: 10694–10699 [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Blesch A, Tuszynski MH 2009. Spinal cord injury: Plasticity, regeneration and the challenge of translational drug development. Trends Neurosci 32: 41–47 [DOI] [PubMed] [Google Scholar]
  14. Bolsover S, Fabes J, Anderson PN 2008. Axonal guidance molecules and the failure of axonal regeneration in the adult mammalian spinal cord. Restor Neurol Neurosci 26: 117–130 [PubMed] [Google Scholar]
  15. Bradbury EJ, Khemani S, Von R, King, Priestley JV, McMahon SB 1999. NT-3 promotes growth of lesioned adult rat sensory axons ascending in the dorsal columns of the spinal cord. Eur J Neurosci 11: 3873–3883 [DOI] [PubMed] [Google Scholar]
  16. Bradbury EJ, Moon LD, Popat RJ, King VR, Bennett GS, Patel PN, Fawcett JW, McMahon SB 2002. Chondroitinase ABC promotes functional recovery after spinal cord injury. Nature 416: 636–640 [DOI] [PubMed] [Google Scholar]
  17. Bregman BS, Kunkel-Bagden E, Schnell L, Dai HN, Gao D, Schwab ME 1995. Recovery from spinal cord injury mediated by antibodies to neurite growth inhibitors. Nature 378: 498–501 [DOI] [PubMed] [Google Scholar]
  18. Bruckner G, Grosche J, Schmidt S, Hartig W, Margolis RU, Delpech B, Seidenbecher CI, Czaniera R, Schachner M 2000. Postnatal development of perineuronal nets in wild-type mice and in a mutant deficient in tenascin-R. J Comp Neurol 428: 616–629 [DOI] [PubMed] [Google Scholar]
  19. Bundesen LQ, Scheel TA, Bregman BS, Kromer LF 2003. Ephrin-B2 and EphB2 regulation of astrocyte-meningeal fibroblast interactions in response to spinal cord lesions in adult rats. J Neurosci 23: 7789–7800 [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Cao Z, Qiu J, Domeniconi M, Hou J, Bryson JB, Mellado W, Filbin MT 2007. The inhibition site on myelin-associated glycoprotein is within Ig-domain 5 and is distinct from the sialic acid binding site. J Neurosci 27: 9146–9154 [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Carmichael ST, Archibeque I, Luke L, Nolan T, Momiy J, Li S 2005. Growth-associated gene expression after stroke: evidence for a growth-promoting region in peri-infarct cortex. Exp Neurol 193: 291–311 [DOI] [PubMed] [Google Scholar]
  22. Caroni P, Schwab ME 1988. Two membrane protein fractions from rat central myelin with inhibitory properties for neurite growth and fibroblast spreading. J Cell Biol 106: 1281–1288 [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Carter LM, Starkey ML, Akrimi SF, Davies M, McMahon SB, Bradbury EJ 2008. The yellow fluorescent protein (YFP-H) mouse reveals neuroprotection as a novel mechanism underlying chondroitinase ABC-mediated repair after spinal cord injury. J Neurosci 28: 14107–14120 [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Chen MS, Huber AB, van der Haar ME, Frank M, Schnell L, Spillmann AA, Christ F, Schwab ME 2000. Nogo-A is a myelin-associated neurite outgrowth inhibitor and an antigen for monoclonal antibody IN-1. Nature 403: 434–439 [DOI] [PubMed] [Google Scholar]
  25. Chivatakarn O, Kaneko S, He Z, Tessier-Lavigne M, Giger RJ 2007. The Nogo-66 receptor NgR1 is required only for the acute growth cone-collapsing but not the chronic growth-inhibitory actions of myelin inhibitors. J Neurosci 27: 7117–7124 [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Corriveau RA, Huh GS, Shatz CJ 1998. Regulation of class I MHC gene expression in the developing and mature CNS by neural activity. Neuron 21: 505–520 [DOI] [PubMed] [Google Scholar]
  27. Costigan M, Befort K, Karchewski L, Griffin RS, D'Urso D, Allchorne A, Sitarski J, Mannion JW, Pratt RE, Woolf CJ 2002. Replicate high-density rat genome oligonucleotide microarrays reveal hundreds of regulated genes in the dorsal root ganglion after peripheral nerve injury. BMC Neurosci 3: 16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Courtine G, Song B, Roy RR, Zhong H, Herrmann JE, Ao Y, Qi J, Edgerton VR, Sofroniew MV 2008. Recovery of supraspinal control of stepping via indirect propriospinal relay connections after spinal cord injury. Nat Med 14: 69–74 [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Curtis R, Adryan KM, Zhu Y, Harkness PJ, Lindsay RM, DiStefano PS 1993. Retrograde axonal transport of ciliary neurotrophic factor is increased by peripheral nerve injury. Nature 365: 253–255 [DOI] [PubMed] [Google Scholar]
  30. Davies SJ, Field PM, Raisman G 1996. Regeneration of cut adult axons fails even in the presence of continuous aligned glial pathways. Exp Neurol 142: 203–216 [DOI] [PubMed] [Google Scholar]
  31. Davies SJ, Goucher DR, Doller C, Silver J 1999. Robust regeneration of adult sensory axons in degenerating white matter of the adult rat spinal cord. J Neurosci 19: 5810–5822 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. DeBellard ME, Tang S, Mukhopadhyay G, Shen YJ, Filbin MT 1996. Myelin-associated glycoprotein inhibits axonal regeneration from a variety of neurons via interaction with a sialoglycoprotein. Mol Cell Neurosci 7: 89–101 [DOI] [PubMed] [Google Scholar]
  33. Dimou L, Schnell L, Montani L, Duncan C, Simonen M, Schneider R, Liebscher T, Gullo M, Schwab ME 2006. Nogo-A-deficient mice reveal strain-dependent differences in axonal regeneration. J Neurosci 26: 5591–5603 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Domeniconi M, Cao Z, Spencer T, Sivasankaran R, Wang K, Nikulina E, Kimura N, Cai H, Deng K, Gao Y, et al. 2002. Myelin-associated glycoprotein interacts with the Nogo66 receptor to inhibit neurite outgrowth. Neuron 35: 283–290 [DOI] [PubMed] [Google Scholar]
  35. Fabes J, Anderson P, Brennan C, Bolsover S 2007. Regeneration-enhancing effects of EphA4 blocking peptide following corticospinal tract injury in adult rat spinal cord. Eur J Neurosci 26: 2496–2505 [DOI] [PubMed] [Google Scholar]
  36. Fabes J, Anderson P, Yanez-Munoz RJ, Thrasher A, Brennan C, Bolsover S 2006. Accumulation of the inhibitory receptor EphA4 may prevent regeneration of corticospinal tract axons following lesion. Eur J Neurosci 23: 1721–1730 [DOI] [PubMed] [Google Scholar]
  37. Figueroa JD, Benton RL, Velazquez I, Torrado AI, Ortiz CM, Hernandez CM, Diaz JJ, Magnuson DS, Whittemore SR, Miranda JD 2006. Inhibition of EphA7 up-regulation after spinal cord injury reduces apoptosis and promotes locomotor recovery. J Neurosci Res 84: 1438–1451 [DOI] [PubMed] [Google Scholar]
  38. Filbin MT 2003. Myelin-associated inhibitors of axonal regeneration in the adult mammalian CNS. Nat Rev Neurosci 4: 703–713 [DOI] [PubMed] [Google Scholar]
  39. Fitch MT, Doller C, Combs CK, Landreth GE, Silver J 1999. 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 19: 8182–8198 [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Fouad K, Schnell L, Bunge MB, Schwab ME, Liebscher T, Pearse DD 2005. Combining Schwann cell bridges and olfactory-ensheathing glia grafts with chondroitinase promotes locomotor recovery after complete transection of the spinal cord. J Neurosci 25: 1169–1178 [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Fournier AE, GrandPre T, Strittmatter SM 2001. Identification of a receptor mediating Nogo-66 inhibition of axonal regeneration. Nature 409: 341–346 [DOI] [PubMed] [Google Scholar]
  42. Fournier AE, Gould GC, Liu BP, Strittmatter SM 2002. Truncated soluble Nogo receptor binds Nogo-66 and blocks inhibition of axon growth by myelin. J Neurosci 22: 8876–8883 [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Fox AN, Zinn K 2005. The heparan sulfate proteoglycan syndecan is an in vivo ligand for the Drosophila LAR receptor tyrosine phosphatase. Curr Biol 15: 1701–1711 [DOI] [PubMed] [Google Scholar]
  44. Fry EJ, Chagnon MJ, Lopez-Vales R, Tremblay ML, David S 2009. Corticospinal tract regeneration after spinal cord injury in receptor protein tyrosine phosphatase sigma deficient mice. Glia 58: 423–433 [DOI] [PubMed] [Google Scholar]
  45. Fu WY, Chen Y, Sahin M, Zhao XS, Shi L, Bikoff JB, Lai KO, Yung WH, Fu AK, Greenberg ME, et al. 2007. Cdk5 regulates EphA4-mediated dendritic spine retraction through an ephexin1-dependent mechanism. Nat Neurosci 10: 67–76 [DOI] [PubMed] [Google Scholar]
  46. Funakoshi H, Frisen J, Barbany G, Timmusk T, Zachrisson O, Verge VM, Persson H 1993. Differential expression of mRNAs for neurotrophins and their receptors after axotomy of the sciatic nerve. J Cell Biol 123: 455–465 [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Galtrey CM, Fawcett JW 2007. The role of chondroitin sulfate proteoglycans in regeneration and plasticity in the central nervous system. Brain Res Rev 54: 1–18 [DOI] [PubMed] [Google Scholar]
  48. Garcia-Alias G, Barkhuysen S, Buckle M, Fawcett JW 2009. Chondroitinase ABC treatment opens a window of opportunity for task-specific rehabilitation. Nat Neurosci 12: 1145–1151 [DOI] [PubMed] [Google Scholar]
  49. Geremia NM, Pettersson LME, Hasmatali JC, Hryciw T, Danielsen N, Schreyer DJ, Verge VMK 2009. Endogenous BDNF regulates induction of intrinsic neuronal growth programs in injured sensory neurons. Exp Neurol (in press) [DOI] [PubMed] [Google Scholar]
  50. Giehl KM, Tetzlaff W 1996. BDNF and NT-3, but not NGF, prevent axotomy-induced death of rat corticospinal neurons in vivo. Eur J Neurosci 8: 1167–1175 [DOI] [PubMed] [Google Scholar]
  51. Giehl KM, Rohrig S, Bonatz H, Gutjahr M, Leiner B, Bartke I, Yan Q, Reichardt LF, Backus C, Welcher AA, et al. 2001. Endogenous brain-derived neurotrophic factor and neurotrophin-3 antagonistically regulate survival of axotomized corticospinal neurons in vivo. J Neurosci 21: 3492–3502 [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Giraudon P, Vincent P, Vuaillat C, Verlaeten O, Cartier L, Marie-Cardine A, Mutin M, Bensussan A, Belin MF, Boumsell L 2004. Semaphorin CD100 from activated T lymphocytes induces process extension collapse in oligodendrocytes and death of immature neural cells. J Immunol 172: 1246–1255 [DOI] [PubMed] [Google Scholar]
  53. Glazner GW, Morrison AE, Ishii DN 1994. Elevated insulin-like growth factor (IGF) gene expression in sciatic nerves during IGF-supported nerve regeneration. Mol Brain Res 25: 265–272 [DOI] [PubMed] [Google Scholar]
  54. Goldshmit Y, Galea MP, Wise G, Bartlett PF, Turnley AM 2004. Axonal regeneration and lack of astrocytic gliosis in EphA4-deficient mice. J Neurosci 24: 10064–10073 [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Gonzenbach RR, Schwab ME 2008. Disinhibition of neurite growth to repair the injured adult CNS: Focusing on Nogo. Cell Mol Life Sci 65: 161–176 [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. GrandPre T, Li S, Strittmatter SM 2002. Nogo-66 receptor antagonist peptide promotes axonal regeneration. Nature 417: 547–551 [DOI] [PubMed] [Google Scholar]
  57. GrandPre T, Nakamura F, Vartanian T, Strittmatter SM 2000. Identification of the Nogo inhibitor of axon regeneration as a Reticulon protein. Nature 403: 439–444 [DOI] [PubMed] [Google Scholar]
  58. Grill R, Murai K, Blesch A, Gage FH, Tuszynski MH 1997. Cellular delivery of neurotrophin-3 promotes corticospinal axonal growth and partial functional recovery after spinal cord injury. J Neurosci 17: 5560–5572 [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Grimpe B, Pressman Y, Lupa MD, Horn KP, Bunge MB, Silver J 2005. The role of proteoglycans in Schwann cell/astrocyte interactions and in regeneration failure at PNS/CNS interfaces. Mol Cell Neurosci 28: 18–29 [DOI] [PubMed] [Google Scholar]
  60. Habib AA, Marton LS, Allwardt B, Gulcher JR, Mikol DD, Hognason T, Chattopadhyay N, Stefansson K 1998. Expression of the oligodendrocyte-myelin glycoprotein by neurons in the mouse central nervous system. J Neurochem 70: 1704–1711 [DOI] [PubMed] [Google Scholar]
  61. Harvey PA, Lee DH, Qian F, Weinreb PH, Frank E 2009. Blockade of Nogo receptor ligands promotes functional regeneration of sensory axons after dorsal root crush. J Neurosci 29: 6285–6295 [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. He XL, Bazan JF, McDermott G, Park JB, Wang K, Tessier-Lavigne M, He Z, Garcia KC 2003. Structure of the Nogo receptor ectodomain: A recognition module implicated in myelin inhibition. Neuron 38: 177–185 [DOI] [PubMed] [Google Scholar]
  63. Höke A, Cheng C, Zochodne DW 2000. Expression of glial cell line-derived neurotrophic factor family of growth factors in peripheral nerve injury in rats. Neuroreport 11: 1651–1654 [DOI] [PubMed] [Google Scholar]
  64. Hollis ER II, Lu P, Blesch A, Tuszynski MH 2009a. IGF-I gene delivery promotes corticospinal neuronal survival but not regeneration after adult CNS injury. Exp Neurol 215: 53–59 [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Hollis ER II, Jamshidi P, Löw K, Blesch A, Tuszynski MH 2009b. Induction of corticospinal regeneration by lentiviral trkB-induced Erk activation. Proc Natl Acad Sci 106: 7215–7220 [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Houle JD, Tom VJ, Mayes D, Wagoner G, Phillips N, Silver J 2006. Combining an autologous peripheral nervous system “bridge” and matrix modification by chondroitinase allows robust, functional regeneration beyond a hemisection lesion of the adult rat spinal cord. J Neurosci 26: 7405–7415 [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Hu F, Strittmatter SM 2008. The N-terminal domain of Nogo-A inhibits cell adhesion and axonal outgrowth by an integrin-specific mechanism. J Neurosci 28: 1262–1269 [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Huh GS, Boulanger LM, Du H, Riquelme PA, Brotz TM, Shatz CJ 2000. Functional requirement for class I MHC in CNS development and plasticity. Science 290: 2155–2159 [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Irizarry-Ramirez M, Willson CA, Cruz-Orengo L, Figueroa J, Velazquez I, Jones H, Foster RD, Whittemore SR, Miranda JD 2005. Upregulation of EphA3 receptor after spinal cord injury. J Neurotrauma 22: 929–935 [DOI] [PubMed] [Google Scholar]
  70. Ji B, Case LC, Liu K, Shao Z, Lee X, Yang Z, Wang J, Tian T, Shulga-Morskaya S, Scott M, et al. 2008. Assessment of functional recovery and axonal sprouting in oligodendrocyte-myelin glycoprotein (OMgp) null mice after spinal cord injury. Mol Cell Neurosci 39: 258–267 [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Jin Y, Fischer I, Tessler A, Houle JD 2002. Transplants of fibroblasts genetically modified to express BDNF promote axonal regeneration from supraspinal neurons following chronic spinal cord injury. Exp Neurol 177: 265–275 [DOI] [PubMed] [Google Scholar]
  72. Johnson PW, Abramow-Newerly W, Seilheimer B, Sadoul R, Tropak MB, Arquint M, Dunn RJ, Schachner M, Roder JC 1989. Recombinant myelin-associated glycoprotein confers neural adhesion and neurite outgrowth function. Neuron 3: 377–385 [DOI] [PubMed] [Google Scholar]
  73. Johnson KG, Tenney AP, Ghose A, Duckworth AM, Higashi ME, Parfitt K, Marcu O, Heslip TR, Marsh JL, Schwarz TL, et al. 2006. The HSPGs Syndecan and Dallylike bind the receptor phosphatase LAR and exert distinct effects on synaptic development. Neuron 49: 517–531 [DOI] [PubMed] [Google Scholar]
  74. Jones LL, Sajed D, Tuszynski MH 2003a. Axonal regeneration through regions of chondroitin sulfate proteoglycan deposition after spinal cord injury: A balance of permissiveness and inhibition. J Neurosci 23: 9276–9288 [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Jones LL, Sajed D, Tuszynski MH 2003b. Axonal regeneration through regions of chondroitin sulfate proteoglycan deposition after spinal cord injury: A balance of permissiveness and inhibition. J Neurosci 23: 9276–9288 [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Jones LL, Yamaguchi Y, Stallcup WB, Tuszynski MH 2002. NG2 is a major chondroitin sulfate proteoglycan produced after spinal cord injury and is expressed by macrophages and oligodendrocyte progenitors. J Neurosci 22: 2792–2803 [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Kadoya K, Tsukada S, Lu P, Coppola G, Geschwind DH, Filbin MT, Blesch A, Tuszynski MH 2009. Combined intrinsic and extrinsic neuronal mechanisms facilitate bridging axonal regeneration one year after spinal cord injury. Neuron 64: 165–172 [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Kaneko S, Iwanami A, Nakamura M, Kishino A, Kikuchi K, Shibata S, Okano HJ, Ikegami T, Moriya A, Konishi O, et al. 2006. A selective Sema3A inhibitor enhances regenerative responses and functional recovery of the injured spinal cord. Nat Med 12: 1380–1389 [DOI] [PubMed] [Google Scholar]
  79. Kikuchi K, Kishino A, Konishi O, Kumagai K, Hosotani N, Saji I, Nakayama C, Kimura T 2003. In vitro and in vivo characterization of a novel semaphorin 3A inhibitor, SM-216289 or xanthofulvin. J Biol Chem 278: 42985–42991 [DOI] [PubMed] [Google Scholar]
  80. Kim JE, Li S, GrandPre T, Qiu D, Strittmatter SM 2003. Axon regeneration in young adult mice lacking Nogo-A/B. Neuron 38: 187–199 [DOI] [PubMed] [Google Scholar]
  81. Kim JE, Liu BP, Park JH, Strittmatter SM 2004. Nogo-66 receptor prevents raphespinal and rubrospinal axon regeneration and limits functional recovery from spinal cord injury. Neuron 44: 439–451 [DOI] [PubMed] [Google Scholar]
  82. Kishino A, Ishige Y, Tatsuno T, Nakayama C, Noguchi H 1997. BDNF prevents and reverses adult rat motor neuron degeneration and induces axonal outgrowth. Experimental Neurology 144: 273–286 [DOI] [PubMed] [Google Scholar]
  83. Klein R 2009. Bidirectional modulation of synaptic functions by Eph/ephrin signaling. Nat Neurosci 12: 15–20 [DOI] [PubMed] [Google Scholar]
  84. Kobayashi NR, Fan D-P, Giehl KM, Bedard AM, Wiegand SJ, Tetzlaff W 1997. BDNF and NT-4/5 prevent atrophy of rat rubrospinal neurons after cervical axotomy, stimulate GAP-43 and Tα 1-tubulin mRNA expression, and promote axonal regeneration. J Neurosci 17: 9583–9595 [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Kury P, Abankwa D, Kruse F, Greiner-Petter R, Muller HW 2004. Gene expression profiling reveals multiple novel intrinsic and extrinsic factors associated with axonal regeneration failure. Eur J Neurosci 19: 32–42 [DOI] [PubMed] [Google Scholar]
  86. Kwon BK, Liu J, Lam C, Plunet W, Oschipok LW, Hauswirth W, Di Polo A, Blesch A, Tetzlaff W 2007. Brain-derived neurotrophic factor gene transfer with adeno-associated viral and lentiviral vectors prevents rubrospinal neuronal atrophy and stimulates regeneration-associated gene expression after acute cervical spinal cord injury. Spine 32: 1164–1173 [DOI] [PubMed] [Google Scholar]
  87. Kwon BK, Liu J, Messerer C, Kobayashi NR, McGraw J, Oschipok L, Tetzlaff W 2002. Survival and regeneration of rubrospinal neurons 1 year after spinal cord injury. Proc Natl Acad Sci 99: 3246–3251 [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Lee JK, Case LC, Chan AF, Zhu Y, Tessier-Lavigne M, Zheng B 2009a. Generation of an OMgp allelic series in mice. Genesis 47: 751–756 [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Lee JK, Chan AF, Luu SM, Zhu Y, Ho C, Tessier-Lavigne M, Zheng B 2009b. Reassessment of corticospinal tract regeneration in Nogo-deficient mice. J Neurosci 29: 8649–8654 [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Lee H, Raiker SJ, Venkatesh K, Geary R, Robak LA, Zhang Y, Yeh HH, Shrager P, Giger RJ 2008. Synaptic function for the Nogo-66 receptor NgR1: Regulation of dendritic spine morphology and activity-dependent synaptic strength. J Neurosci 28: 2753–2765 [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Li L, Hutchins BI, Kalil K 2009. Wnt5a Induces Simultaneous Cortical Axon Outgrowth and Repulsive Axon Guidance through Distinct Signaling Mechanisms. J Neurosci 29: 5873–5883 [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Li S, Liu BP, Budel S, Li M, Ji B, Walus L, Li W, Jirik A, Rabacchi S, Choi E, et al. 2004. Blockade of Nogo-66, myelin-associated glycoprotein, and oligodendrocyte myelin glycoprotein by soluble Nogo-66 receptor promotes axonal sprouting and recovery after spinal injury. J Neurosci 24: 10511–10520 [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Liebl DJ, Morris CJ, Henkemeyer M, Parada LF 2003. mRNA expression of ephrins and Eph receptor tyrosine kinases in the neonatal and adult mouse central nervous system. J Neurosci Res 71: 7–22 [DOI] [PubMed] [Google Scholar]
  94. Liu BP, Fournier A, GrandPre T, Strittmatter SM 2002a. Myelin-associated glycoprotein as a functional ligand for the Nogo-66 receptor. Science 297: 1190–1193 [DOI] [PubMed] [Google Scholar]
  95. Liu Y, Himes BT, Murray M, Tessler A, Fischer I 2002b. Grafts of BDNF-producing fibroblasts rescue axotomized rubrospinal neurons and prevent their atrophy. Exp Neurol 178: 150–164 [DOI] [PubMed] [Google Scholar]
  96. Liu Y, Kim D, Himes BT, Chow SY, Schallert T, Murray M, Tessler A, Fischer I 1999. Transplants of fibroblasts genetically modified to express BDNF promote regeneration of adult rat rubrospinal axons and recovery of forelimb function. J Neurosci 19: 4370–4387 [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Liu Y, Shi J, Lu CC, Wang ZB, Lyuksyutova AI, Song XJ, Zou Y 2005. Ryk-mediated Wnt repulsion regulates posterior-directed growth of corticospinal tract. Nat Neurosci 8: 1151–1159 [DOI] [PubMed] [Google Scholar]
  98. Liu Y, Wang X, Lu CC, Kerman R, Steward O, Xu XM, Zou Y 2008. Repulsive Wnt signaling inhibits axon regeneration after CNS injury. J Neurosci 28: 8376–8382 [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Löw K, Culbertson M, Bradke F, Tessier-Lavigne M, Tuszynski MH 2008. Netrin-1 is a novel myelin-associated inhibitor to axon growth. J Neurosci 28: 1099–1108 [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Low LK, Liu XB, Faulkner RL, Coble J, Cheng HJ 2008. Plexin signaling selectively regulates the stereotyped pruning of corticospinal axons from visual cortex. Proc Natl Acad Sci 105: 8136–8141 [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Lu P, Tuszynski MH 2008. Growth factors and combinatorial therapies for CNS regeneration. Experimental Neurology 209: 313–320 [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Lu P, Blesch A, Tuszynski MH 2001. Neurotrophism without neurotropism: BDNF promotes survival but not growth of lesioned corticospinal neurons. J Comp Neurol 436: 456–470 [DOI] [PubMed] [Google Scholar]
  103. Lu P, Jones LL, Tuszynski MH 2005. BDNF-expressing marrow stromal cells support extensive axonal growth at sites of spinal cord injury. Experimental Neurology 191: 344–360 [DOI] [PubMed] [Google Scholar]
  104. Lu P, Jones LL, Tuszynski MH 2007. Axon regeneration through scars and into sites of chronic spinal cord injury. Exp Neurol 203: 8–21 [DOI] [PubMed] [Google Scholar]
  105. Lu P, Yang H, Jones LL, Filbin MT, Tuszynski MH 2004a. Combinatorial therapy with neurotrophins and cAMP promotes axonal regeneration beyond sites of spinal cord injury. J Neurosci 24: 6402–6409 [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Lu W, Yamamoto V, Ortega B, Baltimore D 2004b. Mammalian Ryk is a Wnt coreceptor required for stimulation of neurite outgrowth. Cell 119: 97–108 [DOI] [PubMed] [Google Scholar]
  107. Lyuksyutova AI, Lu C-C, Milanesio N, King LA, Guo N, Wang Y, Nathans J, Tessier-Lavigne M, Zou Y 2003. Anterior-posterior guidance of commissural axons by Wnt-frizzled signaling. Science 302: 1984–1988 [DOI] [PubMed] [Google Scholar]
  108. Maisonpierre PC, Belluscio L, Friedman B, Alderson RF, Wiegand SJ, Furth ME, Lindsay RM, Yancopoulos GD 1990. NT-3, BDNF, and NGF in the developing rat nervous system: Parallel as well as reciprocal patterns of expression. Neuron 5: 501–509 [DOI] [PubMed] [Google Scholar]
  109. McGee AW, Yang Y, Fischer QS, Daw NW, Strittmatter SM 2005. Experience-driven plasticity of visual cortex limited by myelin and Nogo receptor. Science 309: 2222–2226 [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. McKeon RJ, Hoke A, Silver J 1995. Injury-induced proteoglycans inhibit the potential for laminin-mediated axon growth on astrocytic scars. Exp Neurol 136: 32–43 [DOI] [PubMed] [Google Scholar]
  111. McLean J, Batt J, Doering LC, Rotin D, Bain JR 2002. Enhanced rate of nerve regeneration and directional errors after sciatic nerve injury in receptor protein tyrosine phosphatase sigma knock-out mice. J Neurosci 22: 5481–5491 [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Meyer M, Matsuoka I, Wetmore C, Olson L, Thoenen H 1992. Enhanced synthesis of brain-derived neurotrophic factor in the lesioned peripheral nerve: Different mechanisms are responsible for the regulation of BDNF and NGF mRNA. J Cell Biol 119: 45–54 [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Mi S, Lee X, Shao Z, Thill G, Ji B, Relton J, Levesque M, Allaire N, Perrin S, Sands B, et al. 2004. LINGO-1 is a component of the Nogo-66 receptor/p75 signaling complex. Nat Neurosci 7: 221–228 [DOI] [PubMed] [Google Scholar]
  114. Mikol DD, Stefansson K 1988. A phosphatidylinositol-linked peanut agglutinin-binding glycoprotein in central nervous system myelin and on oligodendrocytes. J Cell Biol 106: 1273–1279 [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Miyashita T, Koda M, Kitajo K, Yamazaki M, Takahashi K, Kikuchi A, Yamashita T 2009. Wnt-Ryk signaling mediates axon growth inhibition and limits functional recovery after spinal cord injury. J Neurotrauma 26: 955–964 [DOI] [PubMed] [Google Scholar]
  116. Mizui M, Kumanogoh A, Kikutani H 2009. Immune semaphorins: Novel features of neural guidance molecules. J Clin Immunol 29: 1–11 [DOI] [PubMed] [Google Scholar]
  117. Montolio M, Messeguer J, Masip I, Guijarro P, Gavin R, Antonio Del Rio J, Messeguer A, Soriano E 2009. A semaphorin 3A inhibitor blocks axonal chemorepulsion and enhances axon regeneration. Chem Biol 16: 691–701 [DOI] [PubMed] [Google Scholar]
  118. Moon LD, Asher RA, Rhodes KE, Fawcett JW 2001. Regeneration of CNS axons back to their target following treatment of adult rat brain with chondroitinase ABC. Nat Neurosci 4: 465–466 [DOI] [PubMed] [Google Scholar]
  119. Moreau-Fauvarque C, Kumanogoh A, Camand E, Jaillard C, Barbin G, Boquet I, Love C, Jones EY, Kikutani H, Lubetzki C, et al. 2003. The transmembrane semaphorin Sema4D/CD100, an inhibitor of axonal growth, is expressed on oligodendrocytes and upregulated after CNS lesion. J Neurosci 23: 9229–9239 [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Moretti S, Procopio A, Lazzarini R, Rippo MR, Testa R, Marra M, Tamagnone L, Catalano A 2008. Semaphorin3A signaling controls Fas (CD95)-mediated apoptosis by promoting Fas translocation into lipid rafts. Blood 111: 2290–2299 [DOI] [PubMed] [Google Scholar]
  121. Morgenstern DA, Asher RA, Fawcett JW 2002. Chondroitin sulphate proteoglycans in the CNS injury response. Prog Brain Res 137: 313–332 [DOI] [PubMed] [Google Scholar]
  122. Morita A, Yamashita N, Sasaki Y, Uchida Y, Nakajima O, Nakamura F, Yagi T, Taniguchi M, Usui H, Katoh-Semba R, et al. 2006. Regulation of dendritic branching and spine maturation by semaphorin3A-Fyn signaling. J Neurosci 26: 2971–2980 [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Naveilhan P, ElShamy WM, Ernfors P 1997. Differential regulation of mRNAs for GDNF and its receptors Ret and GDNFRα after sciatic nerve lesion in the mouse. Eur J Neurosci 9: 1450–1460 [DOI] [PubMed] [Google Scholar]
  124. Neumann S, Bradke F, Tessier-Lavigne M, Basbaum AI 2002. Regeneration of sensory axons within the injured spinal cord induced by intraganglionic cAMP elevation. Neuron 34: 885–893 [DOI] [PubMed] [Google Scholar]
  125. Novikov L, Novikova L, Kellerth JO 1995. Brain-derived neurotrophic factor promotes survival and blocks nitric oxide synthase expression in adult rat spinal motoneurons after ventral root avulsion. Neuroscience Letters 200: 45–48 [DOI] [PubMed] [Google Scholar]
  126. Novikov L, Novikova L, Kellerth JO 1997. Brain-derived neurotrophic factor promotes axonal regeneration and long-term survival of adult rat spinal motoneurons in vivo. Neuroscience 79: 765–774 [DOI] [PubMed] [Google Scholar]
  127. Oertle T, van der Haar ME, Bandtlow CE, Robeva A, Burfeind P, Buss A, Huber AB, Simonen M, Schnell L, Brosamle C, et al. 2003. Nogo-A inhibits neurite outgrowth and cell spreading with three discrete regions. J Neurosci 23: 5393–5406 [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Oudega M, Hagg T 1999. Neurotrophins promote regeneration of sensory axons in the adult rat spinal cord. Brain Res 818: 431–438 [DOI] [PubMed] [Google Scholar]
  129. Paradis S, Harrar DB, Lin Y, Koon AC, Hauser JL, Griffith EC, Zhu L, Brass LF, Chen C, Greenberg ME 2007. An RNAi-based approach identifies molecules required for glutamatergic and GABAergic synapse development. Neuron 53: 217–232 [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Park JB, Yiu G, Kaneko S, Wang J, Chang J, He XL, Garcia KC, He Z 2005. A TNF receptor family member, TROY, is a coreceptor with Nogo receptor in mediating the inhibitory activity of myelin inhibitors. Neuron 45: 345–351 [DOI] [PubMed] [Google Scholar]
  131. Pasterkamp RJ, Giger RJ 2009. Semaphorin function in neural plasticity and disease. Curr Opin Neurobiol 19: 263–274 [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Pasterkamp RJ, Verhaagen J 2006. Semaphorins in axon regeneration: Developmental guidance molecules gone wrong? Philos Trans R Soc Lond B Biol Sci 361: 1499–1511 [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Pasterkamp RJ, Anderson PN, Verhaagen J 2001. Peripheral nerve injury fails to induce growth of lesioned ascending dorsal column axons into spinal cord scar tissue expressing the axon repellent Semaphorin3A. Eur J Neurosci 13: 457–471 [DOI] [PubMed] [Google Scholar]
  134. Pasterkamp RJ, Kolk SM, Hellemons AJ, Kolodkin AL 2007. Expression patterns of semaphorin7A and plexinC1 during rat neural development suggest roles in axon guidance and neuronal migration. BMC Dev Biol 7: 98. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Pearse DD, Pereira FC, Marcillo AE, Bates ML, Berrocal YA, Filbin MT, Bunge MB 2004. cAMP and Schwann cells promote axonal growth and functional recovery after spinal cord injury. Nat Med 10: 610–616 [DOI] [PubMed] [Google Scholar]
  136. Peng X, Zhou Z, Hu J, Fink DJ, Mata M 2009. Soluble nogo receptor down-regulates expression of neuronal nogo-A to enhance axonal regeneration. J Biol Chem. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Pizzorusso T, Medini P, Berardi N, Chierzi S, Fawcett JW, Maffei L 2002. Reactivation of ocular dominance plasticity in the adult visual cortex. Science 298: 1248–1251 [DOI] [PubMed] [Google Scholar]
  138. Prinjha R, Moore SE, Vinson M, Blake S, Morrow R, Christie G, Michalovich D, Simmons DL, Walsh FS 2000. Inhibitor of neurite outgrowth in humans. Nature 403: 383–384 [DOI] [PubMed] [Google Scholar]
  139. Qiu J, Cai D, Dai H, McAtee M, Hoffman PN, Bregman BS, Filbin MT 2002. Spinal Axon Regeneration Induced by Elevation of Cyclic AMP. Neuron 34: 895–903 [DOI] [PubMed] [Google Scholar]
  140. Ramer MS, Priestley JV, McMahon SB 2000. Functional regeneration of sensory axons into the adult spinal cord. Nature 403: 312–316 [DOI] [PubMed] [Google Scholar]
  141. Reza JN, Gavazzi I, Cohen J 1999. Neuropilin-1 is expressed on adult mammalian dorsal root ganglion neurons and mediates semaphorin3a/collapsin-1-induced growth cone collapse by small diameter sensory afferents. Mol Cell Neurosci 14: 317–326 [DOI] [PubMed] [Google Scholar]
  142. Robak LA, Venkatesh K, Lee H, Raiker SJ, Duan Y, Lee-Osbourne J, Hofer T, Mage RG, Rader C, Giger RJ 2009. Molecular basis of the interactions of the Nogo-66 receptor and its homolog NgR2 with myelin-associated glycoprotein: Development of NgROMNI-Fc, a novel antagonist of CNS myelin inhibition. J Neurosci 29: 5768–5783 [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Sapieha PS, Duplan L, Uetani N, Joly S, Tremblay ML, Kennedy TE, Di Polo A 2005. Receptor protein tyrosine phosphatase sigma inhibits axon regrowth in the adult injured CNS. Mol Cell Neurosci 28: 625–635 [DOI] [PubMed] [Google Scholar]
  144. Schnell L, Schwab ME 1990. Axonal regeneration in the rat spinal cord produced by an antibody against myelin-associated neurite growth inhibitors. Nature 343: 269–272 [DOI] [PubMed] [Google Scholar]
  145. Schwab ME 2004. Nogo and axon regeneration. Curr Opin Neurobiol 14: 118–124 [DOI] [PubMed] [Google Scholar]
  146. Sendtner M, Stockli KA, Thoenen H 1992. Synthesis and localization of ciliary neurotrophic factor in the sciatic nerve of the adult rat after lesion and during regeneration. J Cell Biol 118: 139–148 [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Shao Z, Browning JL, Lee X, Scott ML, Shulga-Morskaya S, Allaire N, Thill G, Levesque M, Sah D, McCoy JM, et al. 2005. TAJ/TROY, an orphan TNF receptor family member, binds Nogo-66 receptor 1 and regulates axonal regeneration. Neuron 45: 353–359 [DOI] [PubMed] [Google Scholar]
  148. Shen H, Chung JM, Chung K 1999. Expression of neurotrophin mRNAs in the dorsal root ganglion after spinal nerve injury. Mol Brain Res 64: 186–192 [DOI] [PubMed] [Google Scholar]
  149. Shen Y, Tenney AP, Busch SA, Horn KP, Cuascut FX, Liu K, He Z, Silver J, Flanagan JG 2009. PTP{sigma} Is a receptor for chondroitin sulfate proteoglycan, an inhibitor of neural regeneration. Science 326: 592–596 [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Simonen M, Pedersen V, Weinmann O, Schnell L, Buss A, Ledermann B, Christ F, Sansig G, van der Putten H, Schwab ME 2003. Systemic deletion of the myelin-associated outgrowth inhibitor Nogo-A improves regenerative and plastic responses after spinal cord injury. Neuron 38: 201–211 [DOI] [PubMed] [Google Scholar]
  151. Snow DM, Lemmon V, Carrino DA, Caplan AI, Silver J 1990. Sulfated proteoglycans in astroglial barriers inhibit neurite outgrowth in vitro. Exp Neurol 109: 111–130 [DOI] [PubMed] [Google Scholar]
  152. Sobel RA 2005. Ephrin A receptors and ligands in lesions and normal-appearing white matter in multiple sclerosis. Brain Pathol 15: 35–45 [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Sofroniew MV, Galletly NP, Isacson O, Svendsen CN 1990. Survival of adult basal forebrain cholinergic neurons after loss of target neurons. Science 247: 338–342 [DOI] [PubMed] [Google Scholar]
  154. Song XY, Li F, Zhang FH, Zhong JH, Zhou XF 2008. Peripherally-derived BDNF promotes regeneration of ascending sensory neurons after spinal cord injury. PLoS ONE 3: e1707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Steward O, Sharp K, Yee KM, Hofstadter M 2008. A re-assessment of the effects of a Nogo-66 receptor antagonist on regenerative growth of axons and locomotor recovery after spinal cord injury in mice. Exp Neurol 209: 446–468 [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Steward O, Zheng B, Banos K, Yee KM 2007. Response to: Kim et al. “axon regeneration in young adult mice lacking Nogo-A/B.” Neuron 38: 187–199 Neuron 54:191–195 [DOI] [PubMed] [Google Scholar]
  157. Suzuki K, Kumanogoh A, Kikutani H 2008. Semaphorins and their receptors in immune cell interactions. Nat Immunol 9: 17–23 [DOI] [PubMed] [Google Scholar]
  158. Syken J, Grandpre T, Kanold PO, Shatz CJ 2006. PirB restricts ocular-dominance plasticity in visual cortex. Science 313: 1795–1800 [DOI] [PubMed] [Google Scholar]
  159. Takai T 2005. Paired immunoglobulin-like receptors and their MHC class I recognition. Immunology 115: 433–440 [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Tang S, Shen YJ, DeBellard ME, Mukhopadhyay G, Salzer JL, Crocker PR, Filbin MT 1997. Myelin-associated glycoprotein interacts with neurons via a sialic acid binding site at ARG118 and a distinct neurite inhibition site. J Cell Biol 138: 1355–1366 [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Taylor L, Jones L, Tuszynski MH, Blesch A 2006. Neurotrophin-3 gradients established by lentiviral gene delivery promote short-distance axonal bridging beyond cellular grafts in the injured spinal cord. J Neurosci 26: 9713–9721 [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Tester NJ, Howland DR 2008. Chondroitinase ABC improves basic and skilled locomotion in spinal cord injured cats. Exp Neurol 209: 483–496 [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Thallmair M, Metz GA, Z'Graggen WJ, Raineteau O, Kartje GL, Schwab ME 1998. Neurite growth inhibitors restrict plasticity and functional recovery following corticospinal tract lesions. Nat Neurosci 1: 124–131 [DOI] [PubMed] [Google Scholar]
  164. Thompson KM, Uetani N, Manitt C, Elchebly M, Tremblay ML, Kennedy TE 2003. Receptor protein tyrosine phosphatase sigma inhibits axonal regeneration and the rate of axon extension. Mol Cell Neurosci 23: 681–692 [DOI] [PubMed] [Google Scholar]
  165. Tuszynski MH, Gabriel K, Gage FH, Suhr S, Meyer S, Rosetti A 1996. Nerve growth factor delivery by gene transfer induces differential outgrowth of sensory, motor, and noradrenergic neurites after adult spinal cord injury. Experimental Neurology 137: 157–173 [DOI] [PubMed] [Google Scholar]
  166. Tuszynski MH, Gabriel K, Gerhardt K, Szollar S 1999. Human spinal cord retains substantial structural mass in chronic stages after injury. J Neurotrauma 16: 523–531 [DOI] [PubMed] [Google Scholar]
  167. Venkatesh K, Chivatakarn O, Lee H, Joshi PS, Kantor DB, Newman BA, Mage R, Rader C, Giger RJ 2005. The Nogo-66 receptor homolog NgR2 is a sialic acid-dependent receptor selective for myelin-associated glycoprotein. J Neurosci 25: 808–822 [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Venkatesh K, Chivatakarn O, Sheu SS, Giger RJ 2007. Molecular dissection of the myelin-associated glycoprotein receptor complex reveals cell type-specific mechanisms for neurite outgrowth inhibition. J Cell Biol 177: 393–399 [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. Vinson M, Strijbos PJ, Rowles A, Facci L, Moore SE, Simmons DL, Walsh FS 2001. Myelin-associated glycoprotein interacts with ganglioside GT1b. A mechanism for neurite outgrowth inhibition. J Biol Chem 276: 20280–20285 [DOI] [PubMed] [Google Scholar]
  170. Vyas AA, Patel HV, Fromholt SE, Heffer-Lauc M, Vyas KA, Dang J, Schachner M, Schnaar RL 2002. Gangliosides are functional nerve cell ligands for myelin-associated glycoprotein (MAG), an inhibitor of nerve regeneration. Proc Natl Acad Sci 99: 8412–8417 [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Wang X, Baughman KW, Basso DM, Strittmatter SM 2006. Delayed Nogo receptor therapy improves recovery from spinal cord contusion. Ann Neurol 60: 540–549 [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Wang KC, Kim JA, Sivasankaran R, Segal R, He Z 2002a. P75 interacts with the Nogo receptor as a co-receptor for Nogo, MAG and OMgp. Nature 420: 74–78 [DOI] [PubMed] [Google Scholar]
  173. Wang KC, Koprivica V, Kim JA, Sivasankaran R, Guo Y, Neve RL, He Z 2002b. Oligodendrocyte-myelin glycoprotein is a Nogo receptor ligand that inhibits neurite outgrowth. Nature 417: 941–944 [DOI] [PubMed] [Google Scholar]
  174. Wang B, Zhang N, Qian KX, Geng JG 2005. Conserved molecular players for axon guidance and angiogenesis. Curr Protein Pept Sci 6: 473–478 [DOI] [PubMed] [Google Scholar]
  175. Worter V, Schweigreiter R, Kinzel B, Mueller M, Barske C, Bock G, Frentzel S, Bandtlow CE 2009. Inhibitory activity of myelin-associated glycoprotein on sensory neurons is largely independent of NgR1 and NgR2 and resides within Ig-Like domains 4 and 5. PLoS ONE 4: e5218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Xu NJ, Henkemeyer M 2009. Ephrin-B3 reverse signaling through Grb4 and cytoskeletal regulators mediates axon pruning. Nat Neurosci 12: 268–276 [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Yiu G, He Z 2006. Glial inhibition of CNS axon regeneration. Nat Rev Neurosci 7: 617–627 [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Zhang Y, Dijkhuizen PA, Anderson PN, Lieberman AR, Verhaagen J 1998. NT-3 delivered by an adenoviral vector induces injured dorsal root axons to regenerate into the spinal cord of adult rats. J Neurosci Res 54: 554–562 [DOI] [PubMed] [Google Scholar]
  179. Zhang JY, Luo X-G, Xian CJ, Liu ZH, Zhou XF 2000. Endogenous BDNF is required for myelination and regeneration of injured sciatic nerve in rodents. Eur J Neurosci 12: 4171–4180 [PubMed] [Google Scholar]
  180. Zheng B, Atwal J, Ho C, Case L, He XL, Garcia KC, Steward O, Tessier-Lavigne M 2005. Genetic deletion of the Nogo receptor does not reduce neurite inhibition in vitro or promote corticospinal tract regeneration in vivo. Proc Natl Acad Sci 102: 1205–1210 [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Zheng B, Ho C, Li S, Keirstead H, Steward O, Tessier-Lavigne M 2003. Lack of enhanced spinal regeneration in Nogo-deficient mice. Neuron 38: 213–224 [DOI] [PubMed] [Google Scholar]
  182. Zuo J, Neubauer D, Dyess K, Ferguson TA, Muir D 1998. Degradation of chondroitin sulfate proteoglycan enhances the neurite-promoting potential of spinal cord tissue. Exp Neurol 154: 654–662 [DOI] [PubMed] [Google Scholar]

Articles from Cold Spring Harbor Perspectives in Biology are provided here courtesy of Cold Spring Harbor Laboratory Press

RESOURCES