Open access peer-reviewed chapter

Roles of Semaphorins in Neurodegenerative Diseases

Written By

Sebastian Quintremil, Fernando Medina Ferrer, Javier Puente, María Elsa Pando and María Antonieta Valenzuela

Submitted: 10 August 2018 Reviewed: 15 October 2018 Published: 10 December 2018

DOI: 10.5772/intechopen.82046

From the Edited Volume

Neurons - Dendrites and Axons

Edited by Gonzalo Emiliano Aranda Abreu and María Elena Hernández Aguilar

Chapter metrics overview

3,553 Chapter Downloads

View Full Metrics

Abstract

Semaphorins are secreted and transmembrane proteins that bind plexin/neuropilin or integrin receptors, providing paracrine axonal guidance signals and ultimately leading to a functional and developed neuronal network. Following semaphorin’s initial discovery, their relevance in the central nervous system (CNS) soon intrigued researchers about the possible links between semaphorins, their receptors and signaling mechanisms and different neurodegenerative diseases. Here, we explore the current knowledge of semaphorin’s function and signaling in Alzheimer’s disease (AD), Parkinson’s disease (PD), Charcot-Marie-Tooth disease (CMT), amyotrophic lateral sclerosis (ALS), multiple sclerosis (MS), and Human T-cell lymphotropic virus type 1 (HTLV-1)-associated myelopathy/tropical spastic paraparesis (HAM/TSP). We focus on the effects of the most known semaphorin subclasses 3A and 4D, yet extending our discussion to other semaphorins that have been found involved in specific neuropathologies and the potential effect of semaphorins modulating the immune system in disorders with inflammatory components. Molecular, cellular, and genetic evidences are reviewed, highlighting the relevance of semaphorins on each disease etiology, pathophysiology, and progression. The newly discovered semaphorin functions in neurological disorders even suggest alternative therapies that may be highly valuable in diseases that have no current cure.

Keywords

  • semaphorin
  • neuropilin
  • plexin
  • neuroimmune cross talk
  • neurodegeneration

1. Introduction

Semaphorins (Sema) are a large family of proteins originally discovered as axon guidance signals during development as signals toward proper innervation of targets [1]. Semaphorin function is fundamental during embryonic development, yet they are also largely expressed in the adult brain. In the past decades, an increasing amount of evidence shows that semaphorins participate in refining synaptogenesis, dendritic and axonal exuberance, remodeling of the synaptic network, and even modulating neuronal response to reactive oxygen species and neuronal apoptosis. The association of semaphorins to neuronal function and cell death was soon explored in the context of neurological diseases [2, 3, 4, 5, 6]. Several reports linking alteration of semaphorin function or expression in neuropathologies opened an unexplored door to understand the mechanisms and look for treatment alternatives of disorders with unknown or poorly understood pathological origins. Here, we aim to summarize the state of the art involving semaphorins on Alzheimer’s disease (AD), Parkinson’s disease (PD), Charcot-Marie-tooth disease (CMT), amyotrophic lateral sclerosis (ALS), multiple sclerosis (MS), and Human T-cell lymphotropic virus type 1 (HTLV-1)-associated myelopathy/tropical spastic paraparesis (HAM/TSP caused by HTLV-1 infection).

The role of semaphorins in neurological diseases depends on the different types of semaphorins, their receptors, different signaling pathways they activate, and the neuronal context [1]. For instance, some semaphorins are considered axon repellent in a particular neuronal context/disease, while chemoattractant in others. Similarly, some semaphorins are toxic to neurons and promote apoptosis, while others are neuroprotective. The bifunctionality of semaphorins is mirrored in a myriad of neurological affections and, therefore we, by no means, attempt to provide the reader with a complete list of diseases affected by Sema signaling or a thorough understanding of the Sema family members, their receptor, and their signaling pathways (refer to [2] instead), but rather provide an introductory reading for understanding semaphorin function in neurological pathologies.

1.1 Neurological disorders and the common factor of semaphorin

Neurological disorders of the central nervous system (CNS) are diseases with structural, biochemical, or electrical abnormalities in the brain and spinal cord caused by gene mutations, neuron damage, dysfunction of axon-dendrite connections, myelin loss, and/or damage of the surrounding vascular system. Neurological disorders include AD, PD, CMT, and ALS, sometimes showing an important immune component as MS and HAM/TSP [7, 8]. Currently, semaphorins are related with health and disease in the cardiovascular, immune, and central nervous systems. Although neurological disorders have different pathological origins, the participation of Sema signaling is a common factor [1, 9, 10]. Activation of semaphorin receptors in the neuronal growth cone promotes changes of cytoskeletal dynamics, resulting in an axon extension alteration and therefore possible neuronal dysfunctions in the context of neuropathologies [1, 11].

Advertisement

2. Semaphorin signaling

Semaphorins are a family of eight different subclasses with several members each, grouped based on amino acid sequence and structural similarities. Semaphorins include secreted and membrane-bound glycoproteins that bind mainly to plexin receptors (most relevant semaphorins related with neurological disorders are summarized in Table 1) [1, 2]. Plexins (PLXN) are a family of nine types of transmembrane semaphorin receptors (plexins A1, A2, A3, A4, B1, B2, B3, C1, and D1). Class 3 semaphorins bind to neuropilins (NRP1 and NRP2) that act as co-receptors forming an heterocomplex with type A plexins (the transducing unit), and in some cases, the Sema3-plexin-neuropilin complex may also associate with cell adhesion molecules of the IgCAM superfamily. Class 7 semaphorins bind to integrin receptors instead of plexins and neuropilins [1, 2, 3, 11, 12, 13, 14, 15]. While plexins seem to act as receptors for semaphorins only, the cell surface NRP receptors have pleiotropic functions, being also co-receptors for vascular endothelial growth factor (VEGF). NRP1 has high affinity for VEGF-A and is required for signal transduction after association to the VEGF receptor [16, 17]. Competition between VEGF and Sema3A for partially overlapping binding sites on NRP1 may produce a signaling unbalance potentially involved in neuropathologies.

SemaDiseaseReceptors/coreceptorsCell expressionFunctions
Sema2ACMTPlexin B
Sema1A
NeuronsNeuronal connectivity, cell migration
Sema3A
Sema3B
Sema3C
Sema3D
Sema3E
Sema3F
AD, PD,
ALS, MS
Plexin A1-4/Nrp1/Nrp2, IgCAM, RTK, integrins, proteoglycans VEGFR2Neurons, glia, immune cellsCytoskeletal organization, neuronal connectivity, regeneration, synaptic transmission, regeneration, cell migration, angiogenesis, immune responses
Sema4A
Sema4D
MS, HAM/TSPPlexin B1/Met, ErbB2, Timp2, RTKNeurons, glia, immune cells, endothelial cells, thymusCytoskeletal organization, neuronal connectivity, angiogenesis, cell migration, synaptic transmission, regeneration, immune responses
Sema5AAD, PD,Plexin B3/Nrp2, CSPGs, HSPGs, TKNeurons, gliaCytoskeletal organization, neuronal connectivity, synaptogenesis, vascular patterning
Sema7APD, MSb1-integrins, α,b- integrins, plexin C1Neurons, glia, immune cells, fibroblasts, thymusCytoskeletal organization, cell migration, cell adhesion, immunomodulation, stimulating cytokine production, proinflammatory responses

Table 1.

Semaphorins in neurological diseases.

Receptors [1, 2, 10, 12, 15], cell expression [9, 10] functions [9], and diseases discussed here are indicated.

Once semaphorins bind to their receptors, the transducing unit triggers signaling pathways linking several protein kinases and downstream substrates that overall change microtubule and actin dynamics, promoting growth cone collapse and axon repulsion in neurons. Nevertheless, changes in the neuronal environment or the type of semaphorin/receptor complex may shape a different transduction effect. Plexin receptors contain a cytoplasmic region acting as a GTPase-activating protein to bind and stimulate GTPase activity of Rho, Rac1, Rnd1, and R-Ras proteins. For instance, the G-protein R-Ras is involved in neuronal sprouting and cell adhesion via activation of integrins. Semaphorin signaling via plexin A1/B1 inactivates R-Ras. Sema3A- and Sema4D-mediated signaling, therefore, inhibit integrin β1 subunits through downregulation of R-Ras, leading to a reduction of growth cone adhesion and allowing collapse responses [1, 2, 10, 15]. Another central protein participating in semaphorin-induced growth cone collapse signaling is collapsin response mediator protein-2 (CRMP-2). CRMP-2 is a phosphoprotein mostly expressed in the CNS and involved in the cytoskeleton structure and function of neuronal cells through the induction of microtubule dynamics/assembly by binding to α- and β-tubulin heterodimers. The complex CRMP-2/kinesin-1 regulates soluble tubulin transport to the distal part of the growing axon and also neurite formation by modulating tubulin GTPase via intramolecular interaction with its N-terminal inhibitory region [18, 19, 20]. The affinity of CRMP-2 for tubulin is significantly diminished when specific residues are phosphorylated, leading to axon retraction and growth cone collapse [21, 22].

Advertisement

3. Semaphorin regulatory functions on neuronal and non-neuronal cells

3.1 Axonal degeneration associated to cytoskeleton organization

The roles of Sema3A and Sema4D are to produce alteration of both actin and microtubule dynamics in the cytoskeleton organization (ratio of the polymerization/depolymerization rates). The effects of Sema3A and Sema4D on actin dynamics include the downregulation of PI3K-Akt signaling pathway, inhibiting integrin-mediated adhesion as well as repulsive effects on axonal growing associated to actin-rich structure loss of lamellipodia and filopodia as part of the cofilin pathway triggered by Sema3A, or activation of myosin II (MyoII) and F-actin bundles promoted by Sema4D-ROCK signaling [2, 23, 24]. Sema3A additionally mediates an increase of the nonphosphorylated active form of myosin II (MyoII) and decreases the phosphorylation levels of Ezrin, Radixin, and Moesin (ERM) proteins. The nonphosphorylated active form of MyoII promotes retraction, while low ERM phosphorylation reduces the crosslinking between actin filaments and the plasma membrane [2, 25, 26]. Regarding microtubule dynamics, Sema4D signaling induces the inactivation of CRMP-2 by glycogen synthase kinase 3 beta (GSK3B)-mediated phosphorylation, leading to reduced binding of CRMP-2 to tubulin and consequently limiting its stabilizing function at the plus end of microtubules. Sema3A induces a similar CRMP-2-inactivating mechanism, yet requiring phosphorylation by cyclin-dependent kinase 5 (Cdk5) at Ser522 and GSK3B at Thr509/514 [2, 21, 22].

3.2 Axonal degeneration associated to synaptogenesis and synaptic plasticity

Synapsis formation of the neuronal network requires local participation of cell adhesion molecules, extracellular proteins, and axon guidance molecules at axodendritic sites [6]. Sema3A and VEGF signaling have been proposed to actively modulate the synaptogenesis and synaptic plasticity, since they are dysregulated in neurological diseases, such as AD [4, 16, 17]. Proper regulation of the synapse formation and dendritic branching contributes to a normal balance between excitatory and inhibitory synaptic transmission; dysregulation of this balance would interfere with the regeneration of damaged CNS axons [6, 11, 15, 27, 28].

3.3 Semaphorin function on axonal regeneration

Models of axonal regeneration have been studied after spinal cord injury, which can produce a permanent damage because axonal growth and regeneration are limited after injury [29]. In adult mammals, some myelin proteins produced by myelinating oligodendrocytes, such as Nogo-A, MAG and OMgp, inhibit spinal cord regeneration [4]. Additionally, glial cells release chondroitin sulfate proteoglycans (CSPGs) and lecticans, such as neurocan, brevican, phosphacan, and tenascin to form the neuronal extracellular matrix (perineuronal nets). The perineuronal nets, together with other extracellular proteins, such as, ephrins, slits, netrins, bone morphogenic proteins, Wnts, and semaphorins are among the molecules most likely involved in limiting axonal regeneration [4, 30, 31, 32, 33]. However, the function of semaphorins is dependent on the cellular context and may also favor axon regeneration. For instance, nerve growth factor (NGF) co-injected with Sema3A in trigeminal neuronal cell culture induced neuron regeneration [34]. Sema3A has also been implicated in the restoration of functionally motor innervation required to regenerate fibers [35]. Sema4D has shown enhanced locomotor recovery and axon regeneration when expressed in motoneurons, attributed to regulation of microglia function following spinal cord injury in adult zebrafish [36].

3.4 Semaphorin function on revascularization

Revascularization (restoration of perfusion) is regulated by several growth factors secreted from endothelial cells, such as VEGFA, FGF, and PDGFs [4]. Class 3 semaphorins are considered anti-angiogenic semaphorins (e.g., Sema3A, Sema3E, and Sema3F), because they interfere with the effect of VEGF by competing for the same NRP receptor [37, 38]. Sema3A not only targets the actin cytoskeleton, but also the assembly/disassembly of focal adhesions, altering migration, proliferation, and adherence of endothelial cells [37]. Sema3A and Sema4D also produce alteration on the blood-brain barrier (BBB) by disrupting endothelial tight junctions and thus increasing its permeability. BBB damage has been related to higher infiltration of immune cells mediated by increasing levels of Sema4D in MS [4, 39]. As opposed to Sema3A, Sema4D is pro-angiogenic and promotes endothelial cell migration via plexin-B1-PI3K-Akt signaling pathway [37, 40].

3.5 Semaphorin function on remyelination

Axon myelination in the CNS is essential for regulating fast and slow axonal transport rates. Myelination requires interaction among axons, oligodendrocytes, and semaphorins. Semaphorins regulate the migration of oligodendrocyte precursor cells (OPC) during normal development and toward demyelinated lesions. Demyelinization, caused by loss of oligodendrocytes and myelin sheaths around axons, is a pathological condition that results in axonal dysfunction, degeneration and loss of sensory and motors neurons [2, 4, 41, 42, 43]. Oligodendrocyte death can be produced by genetic defects, infections, autoimmune reactions, and trauma, along with unknown causes. In some CNS pathologies related to myelination, astrocytes clear off myelin debris, modulating oligodendrocyte activity, myelin maintenance, and its synthesis [43]. Semaphorins (Sema3A, Sema4D, Sema5A, Sema6A, and Sema7A) inhibit OPC recruitment into demyelinated lesions and its differentiation to oligodendrocytes [9, 42, 44, 45]. Sema4D, Sema6, and Sema7A have been detected in myelin, and their expression found strongly upregulated in oligodendrocytes located near the injury site [9, 2, 45].

3.6 Semaphorin function on immune responses

Sema4D, formerly known as CD100, was called the “immune semaphorin” because it was originally found in lymphocytes [3]. The Sema4D receptor in neuronal cells is plexin B1, whereas in immune cells, besides binding to plexin B1, Sema4D also binds to a signaling surface receptor CD72. CD72 is considered a regulatory receptor, because it activates suppressive signals and prevents some forms of autoimmunity [46, 47]. Sema4D is a membrane-bound protein that can be proteolytically cleaved by MT1-MMP metalloprotease, releasing a 120-kDa soluble form of Sema4D, which can act paracrinally on other systems [40, 47]. Immune semaphorins also include Sema3A, Sema4A, Sema6D, and Sema7A expressed on T-cells, B-cells, natural killer cells, neutrophils, platelets, and mature dendritic cells (DC) [39, 47]. The neuronal system sense changes to maintain CNS homeostasis and communicates to the immune system by soluble factors to inhibit further inflammatory responses. For instance, neurons control T-cell and glia functions mediated by membrane-bound or secreted molecules such as semaphorins, neurotrophins, neurotransmitters, neuropeptides, and cytokines [48, 49]. Sema3A and Sema7A expression in neurons attenuate T-cell activation, proliferation, and function through T-cell receptor signaling [49, 50, 51]. Sema3A, additionally, downregulates autoimmunity by suppressing B- and T-cell-mediated autoimmune over-activity responses [52]. In addition, cell adhesion molecules expressed by neurons, such as NCAM, cadherins, and integrins are active molecules in neurogenesis and synaptic plasticity that also can ameliorate inflammation and neurotoxic effects, while strengthening neuroprotection of immune components in pathology [48].

Advertisement

4. Semaphorins and their current association to neurological diseases

4.1 Alzheimer’s disease

Alzheimer’s disease (AD) is the most invalidating, common, and widespread elderly associated neuropathology. An estimated of over 30 million people are affected worldwide, increasing its incidence with age [53]. Patients with AD suffer progressive neuron lost, mostly from the prefrontal cortex and the hippocampus. As a consequence of neuronal death, patients experience memory, cognitive, and behavioral problems, leading within an average of less than 10 years to dementia and/or death [53]. Given its prevalence and the continuing aging of the population, AD threatens to generate an epidemic healthcare crisis in the next decades and yet its cause remains unknown. Current treatments target late symptoms, improving the patient’s quality of life; however, they have a minute contribution on the disease impact and its inevitable progression. Two major features are distinct from Alzheimer’s brains that have intrigued researchers since Alois Alzheimer described them in 1906: presenile plaques and neurofibrillary tangles (NFT). Plaques are extracellular insoluble aggregates, mostly composed of a misfolded amyloid beta (Aβ) peptide, whereas NFT are intraneuronal aggregates of hyperphosphorylated Tau (a microtubule-associated protein). It is yet controversial whether tangles, plaques, or both are causes or consequences of AD. Semaphorins have been long suggested to play a role in AD, because they initially were found largely expressed in adult brain tissues [54] and later found to have abnormal neurohistological patterns in affected cortical and hippocampal areas of AD brains [55], especially on vulnerable neurons [56, 57].

It has been hypothesized that Sema3A may be involved in the early stages of Alzheimer’s degeneration. By analyzing the histological localization of Sema3A in vulnerable hippocampal areas that are first affected by neurodegeneration, a dysregulation in Sema3A expression and release has been proposed as a possible early sign in AD brains, observed even in neurons lacking NFT and therefore possible preceding Tau hyperphosphorylation and aggregation [57]. An intracellular form of Sema3A was also found associated to NFT, along with Plexins, hyperphosphorylated CRMP-2, and microtubule-associated protein 1B (MAP-1B) [57]. The association between Sema3A and a possible downstream Tau hyperphosphorylation may come from the kinases involved in Sema3A signaling. Yoshida et al. initially found a phosphorylated form of CRMP-2 associated to NFT and significantly higher in AD brains [58]. CRMP-2 was originally discovered form its chicken homolog, CRMP-62, as a downstream effector of semaphorin (formerly known as collapsin) signaling. Injection of blocking antibodies against CRMP-62 into dorsal root ganglion inhibited the growth cone collapse induced by Sema3A, suggesting that CRMP-62 is a downstream effector in Sema3A signaling [59]. Since then, CRMP-2 has been associated to NFT in AD [58] and found phosphorylated as a result of semaphorin signaling [60, 61]. The signaling mechanism involves phosphorylation of CRMP-2 by GSK3B after a priming phosphorylation at Ser522 [61, 62]. The phosphorylation at Ser522 is required for further phosphorylation of GSK3B. In rat models, Cdk5 has been found responsible of Ser522 phosphorylation, priming the phosphorylation site of GSK3B both in vitro and in vivo [63, 64]. The sequential phosphorylation of CRMP-2 reduces its affinity to tubulin and triggers microtubule destabilization and Sema3A-mediated growth cone collapse associated to AD pathogenesis [63, 64]. Other kinases, such as Fyn, cyclin-dependent kinase 1, and dual specificity tyrosine-phosphorylation-regulated kinase 2 may also be involved in the priming phosphorylation at Ser522 [61, 64, 65]. The participating kinases and the Sema3A-mediated sequential phosphorylation mechanisms of CRMP-2 are strikingly similar to the pathway leading to Tau hyperphosphorylation, aggregation, and microtubule destabilization observed in AD [66].

A recent study also links Tau phosphorylation to Sema3A signaling by discovering several single-nucleotide polymorphisms (SNPs) associated to AD in the PLXNA4 gene, which codifies for plexin A4 [62]. Most of the top-ranked SNPs associated to AD were located in the region coding the Sema3A-binding site [62]. Cells expressing PLXNA4 and stimulated with Sema3A showed Sema3A-induced phosphorylation of Tau, enhanced by overexpression of the full-length PLXNA4 receptor, whereas expression of soluble forms of PLXNA4 inhibited Tau phosphorylation, presumably by binding Sema3A and competing with the endogenous Plexin receptor, both in a cell line model and in rat hippocampal neurons [62]. The full-length PLXNA4 expression was found higher in AD brains, and also significantly correlates with clinical and neuropathological disease severity measures, such as dementia. The Sema3A-induced kinase activities affecting CRMP-2 and Tau may ultimately lead to neurofibrillary tangle formation and neuronal dead in AD. If Sema3A signaling is effectively involved in the early stages of neurodegeneration, it would be worth to further study its association with AD toward the discovery of new biomarkers and drugs, such as specific kinase inhibitors (some of them already in clinical trial) [67]. An example of a different treatment approach involves modulating the interaction of semaphorins with the neuronal extracellular matrix or perineuronal nets. Differential expression of several proteins related to the extracellular matrix, among them Sema3C, has been found in AD-vulnerable brain areas [56]. Additionally, memory restoration in AD mice models has been achieved by digesting CSPs, a major component of perineuronal nets, using chondroitinase [68, 69]. Disruption of perineuronal nets presumably allows the formation of new synapsis sites and thus increases adult brain plasticity. An important effector of perineuronal nets is Sema3A by binding to chondroitin sulfate, a main component of CSPs [70]. A recent study showed restoration of object recognition memory in a tauopathy mice model via reducing Sema3A binding to perineuronal nets by perirhinal cortex injections of an inhibitory proteoglycan-neutralizing antibody against chondroitin 4-sulfate [69]. Therefore, blocking the binding of Sema3A to perineuronal nets can restore memory function in adult AD-mice model.

It is also interesting to note the bifunctional effect of semaphorins in AD. For instance, Sema3A has been shown to promote apoptosis and neurodegeneration [57, 61], whereas Sema3C has been related to neuroprotection [56, 71]. Such duality, along with several other genetic, environmental, and aging components, gives to AD its multifactorial category. Genetic factors are thought to account for over 50% of the disease, yet these risk factors are not determinants to causing AD. In rare cases of early-onset familial AD, the disease is linked to mutations on genes involving Aβ metabolism, such as the amyloid precursor protein (APP) and presenilin (PSEN1/2) genes. APP is a transmembrane protein from which Aβ peptide is generated by the cleavage of a gamma secretase complex. Presenilin-1 and presenilin-2 are part of the gamma secretase complex. However, in the most common sporadic (nonfamilial) late-onset AD, the genetic variants only explain part of the disease etiology. Several genes have been considered a risk factor, such as the APOE-ε4 polymorphic isoform of the apolipoprotein E gene (APOE), the main cholesterol carrier in the CNS. The role of apolipoprotein E in AD is, however, poorly understood and thought to be related to Aβ degradation [72]. Semaphorin polymorphs have also been studied given their relevance in neuronal apoptosis and the findings of semaphorin polymorphisms related to other syndromes. The first attempts in relating semaphorin SNPs with AD were, however, unfruitful. Evaluation of two exonic SNPs of semaphorin 3A and 4D genes in patients with AD showed no correlation, even though modeling analysis predicted a damaged variant of the affected proteins [73]. Recently, however, a novel proposed method for detecting hidden SNPs that would otherwise appear undetected by commonly used tests found six SNPs on noncoding regions near the semaphorin 5A gene [74]. Schott et al. [75] also found in a noncodifying region of the SEMA3C gene a polymorphism associated to posterior cortical atrophy, a variant of AD. The SEMA3C SNP, therefore, suggests a role of Sema3C during development that may influence later neurodegeneration associated to specific brain regions that ultimately lead to differential degeneration phenotypes [75]. In light of these recent associations of semaphorins to AD [62, 74, 75], semaphorin gene variants and their receptors are expected to participate as a risk factor of AD and its associated neuropathologies, opening a relatively underexplored door toward new discoveries.

4.2 Parkinson’s disease

Parkinson’s disease (PD) is the second most prevalent neurodegenerative disorder after AD, affecting up to 1% of the worldwide population above 60 years old [76]. PD patients suffer a progressive and selective loss of dopaminergic neurons in the Substantia Nigra pars compacta (SNpc), which innervates neurons in the dorsal striatum and regulates its effects on motor functions. The result of neurodegeneration manifests mostly with movement disorders or dyskinesia, where tremor at rest, rigidity, and bradykinesia are cardinal for diagnosis [76]. The reason why SNpc dopaminergic neurons die is not well understood. A composite interaction between genetics and environmental factors in the context of aging has been intensively studied to find the causes, preventing strategies and potential therapies of PD. Semaphorins in their role of apoptotic mediators in neurodegeneration have early on been suggested to be involved in the pathogenesis of PD [77, 78, 79, 80].

Even though it is currently accepted that genetic plays a minor role in sporadic (nonfamilial) PD based on studies with relatives [81], several genes are known to be involved in the disease etiology and progression. For instance, mutations in the SNCA gene are known to be associated with familial PD and increased risk of sporadic PD. The SNCA gene encodes alpha-synuclein, which is the main protein found in Lewy bodies as insoluble aggregates inside SNpc neurons of patients with PD. Interestingly, population genetic studies have also found SNPs in the Sema5A gene that may be related to PD. In the pioneer work of Maraganore et al., 11 SNPs were associated to PD in Caucasian Americans by using a two-stage whole-genome association analysis including 198,345 SNPs. The SNP with the lowest p-value associated to increased risk of PD was located in the Sema5A gene (rs7702187 polymorphism, corresponding to a thymine substituted by adenine in an intronic sequence) [82]. Given the relevance of semaphorins in neuronal apoptosis, the polymorphism in Sema5A found by Maraganore et al. [82] was soon assessed by different research groups, though having conflicting results that still remain unclear. The controversies may come from the different populations evaluated in each analysis. Some studies have shown that the rs7702187 polymorphism is not associated to PD [83, 84] or even conflict to the extent of finding the polymorphic variant protective rather than a risk factor on a population-dependent basis [85]. Taiwanese Asians showed significant associations of the Sema5A rs7702187 polymorphism [85], whereas Finnish Caucasians, Polish Caucasians, Singapore Asians, and Chinese Han populations were not associated to PD [83, 84, 85, 86]. In addition, a different SNP, the rs3798097 (C > T), located in the 5’UTR of the Sema5A gene, was also found associated with PD [85]. Ding et al. [86] showed that although the genotypes are not necessarily associated to PD in a Chinese Han population, the haplotypes involving these two SNPs in the context of a particular ethnicity may be implicated in the disease by finding the AC haplotype associated with an increased risk of PD compared to the common TC haplotype, whereas the AT haplotype was found protective against PD [86]. A more recent meta-analysis on the rs7702187 polymorphism concluded that the A allele frequency was associated to increased risk of PD only in Western population [87]. Both polymorphisms do not affect the sequence of the protein, but rather may be regulatory sequences affecting expression dynamics of Sema5A. These SNPs may be associated to PD; however, less studied polymorphic variants together with other cellular and molecular factors should be considered as well, such as the expression of other semaphorin classes and their receptors that may result as a confounding factor across different population genetic studies.

Besides the genetic highlights on PD, the mechanisms by which semaphorins participate in the disease etiology and progression are poorly understood. Although semaphorins had been suggested to promote neurodegeneration in PD, limited studies were known to link semaphorins in the context of the disease pathogenesis [77, 78, 88, 89]. After a decade of research since the study of Maraganore et al. [82], several lines of evidence indicate possible direct mechanisms of semaphorins (though, not Sema5A) in PD etiology. Recent evidence in a neurotoxin (MPTP)-induced PD mouse model directly involves Sema3A effects through Rho-ROCK signaling pathway. Rho kinase inhibition as well as heterozygous mice knockouts in both Rho and ROCK protect from MPTP-induced damage of dopaminergic neurons, increase dopamine and its metabolic products at the striatum, showed reduced protein expression of Sema3A and its receptors, plexin A and NRP-1, and overall alleviate the behavior damage compared to control PD mice [90, 91, 92, 93]. The effects of Sema3A on Rho-ROCK signaling pathway could mediate several cytoskeleton effectors, contributing to growth cone collapse as well as regulation of neuronal apoptosis. Animal models and SNpc of human brains from PD-affected subjects show apparent neuronal apoptotic processes, probably triggered by oxidative stress [94]. Whether semaphorins are directly involved in causing neuronal apoptosis in PD is debatable and may be associated in part to their specific receptors and signaling pathways. For instance, although Sema3A through plexin A/NRP-1 and activation of Rho kinase was found to promote dopaminergic damage [92], Sema7A has been found protective against ROS-mediated neurodegeneration [95]. Sema7A binds to integrin receptors and could potentially regulate apoptosis through different mechanism. Sema7A reduces the large axonal arborization on dopaminergic neurons, which potentially decreases mitochondrial oxygen demand, ROS production, and neuronal vulnerability observed in PD [95]. However, even different ligands to the same receptor for Sema3A have been found protective against neuronal apoptosis. For instance, VEGF shows neuroprotection in PD models likely by binding to neuropilin receptors, though it is unclear whether the downstream VEGF signaling effectors differ from the semaphorin transduction pathway or VEGF indirectly promotes protection by competition to the same receptor. Alternatively, VEGF could mediate apoptosis by other processes such angiogenesis [89]. It would be interesting to evaluate in future research different ROS-mediated apoptotic pathways and their interplay with semaphorin-induced signaling, such as the phosphorylation targets of ROCK resulting in growth cone collapse (which could mediate CRMP-2 phosphorylation by a similar mechanism to what described previously for AD), to find out how these pathways are regulated in an effort to evaluate new drug candidates for PD prevention and treatment.

Paradoxically, the Sema3A found to promote neurodegeneration in PD pathogenesis, at the same time may be useful for steam cell transplantation therapy in PD patients [96, 97, 98, 99]. Given that semaphorins participate in the formation of the nigrostriatal pathway during prenatal development, they have also been proposed to guide axons to their appropriate targets after possible cell replacement therapy with dopaminergic neurons [96, 100, 101, 102, 103, 104, 105]. Embryonic stem cells differentiated to tyrosine hydroxylase-expressing neurons have been shown to have similar phenotype, expression of neuropilins, and response to Class 3 semaphorins than embryonic ventral mesencephalon neurons [96, 97, 106]. Via neuropilin-mediated signaling, Sema3A increases axonal length in collagen gel coculture experiments. Sema3C, besides increasing length, also attracts axons, whereas Sema3F produces either no effect or axon repulsion [96, 97]. Semaphorin axonal guidance results are promising toward the recovery of parkinsonian symptoms in transplanted PD animal models [98, 99]. Therefore, even though semaphorins may be directly involved in promoting PD neurodegeneration, they could also be a strategy to restore the dopaminergic function by providing axon guidance cues after embryonic stem cell intranigral transplantation.

4.3 Charcot-Marie-Tooth disease

Charcot-Marie-Tooth disease (CMT) is an inherited peripheral neuropathy associated with mutations in more than 90 different genes. CMT is divided into different forms based on the inheritance pattern and neurophysiological observations. The most common types are autosomal-dominant forms, and they are categorized into demyelinating with reduced nerve conduction velocities (CMT type 1) and axonal-loss type with relatively normal nerve conduction velocities (CMT type 2). Patients with CMT type 2 comprise about 20% of all cases [107, 108, 109, 110].

Mutations in the gene GARS, encoding glycyl-tRNA synthetase (GlyRS), have been related to peripheral nerve degeneration and CMT type 2 [111]. In addition, mutated GlyRS has shown to bind neuropilin-1 in mice [112]. Besides its housekeeping intracellular function during protein synthesis, GlyRS can be secreted and produce different cellular effects from the extracellular space [113]. In Drosophila, the Cader lab showed that mutant GlyRS is secreted by muscles and interacts with the neuromuscular junction [114]. Recently, they showed that the P234KY mutant version of GlyRS (mutation associated to CMT type 2) colocalizes with plexin B in presynaptic neurons. Also, Sema2A overexpression, but not Sema1A overexpression, decreased the effect that mutant GlyRS produced on muscle contraction, suggesting that plexin B signaling could be affected by mutated GlyRS by competition with Sema2A [115]. Also, other ligands for neuropilin should be taken into account, such as VEGF. He et al. suggest that CMT type 2 mutations in GlyRS promote its abnormal binding to neuropilin-1, antagonizing the binding of VEGF and blocking the VEGF/neuropilin-1 signaling essential for survival and function of motor neurons [112]. Nevertheless, the neuropilin sequestration by mutant GlyRS has shown to be less detrimental in other tissues, given that this abnormal interaction is permissive to maturation and maintenance of the vasculature in CMT type 2 mice [116].

It is important to consider that in addition to the extracellular function, mutated GlyRS can have abnormal intracellular functions that could also contribute to the CMT pathogenesis, suggesting that multiple mechanisms could be participating. For example, human GlyRS mutations related to CMT (S581 L and G598A E71G, L129P, S211F, G240R, E279D, H418R, and G526R) have shown to have a gain-of-function effect binding to histone deacetylase 6 (HDAC6) and enhance its function, promoting α-tubulin de-acetylation and leading to axonal transport deficit. It is relevant to highlight that G598A patients have more severe distal weakness and wasting in the lower limbs, and in that same article, this mutation showed one of the strongest affinities for HDAC6 [117]. Thus, the most severe mutations in GlyRS could eventually promote abnormal interaction with both NRP1 and HDAC6. A combination of intracellular and extracellular effects could eventually explain the severity and early-onset clinical symptoms of the patients carrying the G598A mutation, as the authors suggested. Future experiments will have to address in more detail the contributions that different plexin/neuropilin ligands may have in CMT, and also link the phenotypes with abnormal activation or deactivation of transduction pathways controlled by these receptors.

4.4 Amyotrophic lateral sclerosis

ALS is a neurological disorder with motor neuron degeneration. Neuron loss leads to paralysis in muscles and death mostly by respiratory failure. Most of the studies in animal models related to ALS use superoxide dismutase (SOD) mutations in mice (in particular, the SOD1G93A transgenic mouse), although the mechanism by which SOD mutations cause ALS is not clear. In these mice models, modifications in axons and nerve terminals are observed even before the clinical symptoms [118].

The first report linking semaphorins and ALS was published in 2006 by De Winter et al. showing increased Sema3A mRNA levels in the SOD1G93A transgenic mice model [119]. Nevertheless, a more recent report from the same lab showed that ALS mice expressing a mutant version of Sema3A (K108 N mutation that produces diminished signaling capacity) had no difference in ALS-induced decline in motor behavior, contrary to what was initially expected [120]. These last results point to a minor contribution of Sema3A to ALS pathology, although other articles are in clear contradiction with this claim. A clear example is the article published by Venkova et al. who hypothesized that Sema3A is able to trigger distal axonopathy and muscle denervation in the SOD1G93A mouse model of ALS [121]. They propose that Sema3A released from terminal Schwann cells activates plexin-A/neuropilin-1, promoting the regulation of kinases such as CDK5 and GSK3B that could alter CRMP-2 phosphorylation and leading to microtubule instability and actin cytoskeletal rearrangements. The Sema3A-mediated signaling could inhibit compensatory axon sprouting and coordination of neuromuscular junction remodeling after injury, contributing to distal axonopathy [121]. Anti-neuropilin-1 antibodies that block the Sema3A docking site in differentiated motor neuron-like cells (NSC-34) prevented Sema3A-induced growth cone collapse. Furthermore, injections of blocking antibodies delayed and even temporarily reversed the motor functional decline while prolonging the life span of SOD1G93A mice. Histologically, the antibody reduced neuromuscular junction denervation and attenuated pathologic alterations in ventral roots at late stages of the disease [121] [121].

In parallel, Miyazaki et al. focused on extracellular protein changes in SOD1G93A mice during the development of ALS to characterize changes in the cellular environment that could affect regeneration [122]. They found decreased Sema3A levels in the anterior half of the lumbar cord of ALS mice. Sema3A immunochemistry at ages 15 and 18 weeks showed a progressive decrease of staining in the neuropil of ALS mice compared to wild type, while Sema3A-positive astrocyte appeared [122]. In addition, it was found that Sema3D gene expression levels are decreased 2.5-fold with respect to wild type in another ALS mouse model (SOD1G37R mutation) [123].

Another piece of evidence for the role of semaphorins on ALS is related to microribonucleic acids (miRNAs). miRNAs are small single-stranded, noncoding RNAs that alter gene expression through post-transcriptional regulation by binding to the 3′-untranslated region of target mRNAs [124]. The Perlson lab [125] analyzed miRNA profiles from axons and somas of two ALS mouse models, SOD1 with G93A mutation and TDP43 with A315T mutation. They showed that different miRNAs were significantly altered in the axons expressing ALS mutations, but not in the somas, indicating that miRNA could be regulating local functions in motor neuron axons [125]. Later, the same lab using qRT-PCR showed that one of these miRNAs, miR126-5p, downregulates Sema3A, Sema3B, neuropilin-1, and neuropilin-2 transcript levels in HeLa cells. Primary myoblasts with the SOD1G93A mutation were transfected with miR126-5p and cultured in a distal compartment of a microfluidic chamber together with a motor neuron explant placed in the proximal compartment. They showed that in the microfluidic chamber, the rate of axons that traversed the distal compartment was increased respect to the control condition of myoblasts transfected with an irrelevant miRNA. In addition, the injection of miR126-5p to ALS mice increased the amount of intact neuromuscular junctions revealing higher innervation in treated muscles compared to the mock condition. Three parameters: Mean Stand Index (measurement of the speed at which the paws detach from the walking surface), single-support parameter (the relative duration of all combined paws in contact with the glass floor), and base of support parameter (the average width of limb spreading between front or hind paws) were measured in ALS mice, and in all cases, the injection of miR126-5p improved all parameters respect to the control [126]. Based on these observations and previous reports, the authors suggested an attractive model of Sema3A/neuropilin-1 interaction that explains how the motor neuron degeneration in ALS could be regulated by miR126-5p. miR126-5p decrease in ALS could enhance Sema3A secretion in muscle and overexpression of neuropilin-1 in axons, increasing Sema3A signaling in the neuromuscular junction and leading to axon degeneration [126].

It is of consideration to test the results obtained with ALS mouse models in human samples. Motor cortex tissue samples showed increased Sema3A mRNA levels by quantitative RT-PCR in ALS patients (eight cases aged 44–72 years) compared to control samples (six subjects aged 45–84 years, with no neurological disease history). Likewise, by immunohistochemistry, the motor cortex showed stronger cytoplasmic and axonal Sema3A labeling in motor neurons of ALS patients compared to controls. Sema3A mRNA levels and immunohistological labeling showed, however, no difference between ALS patients and controls in spinal cord tissue samples [127]. Sema3A levels in human samples support the previous findings in ALS mouse models discussed above. However, other semaphorins and neurological factors not studied yet in the context of ALS may provide a better understanding of semaphorin function and mechanisms on ALS pathology.

4.5 Spastic paraparesis associated to HTLV-1

HTLV-1-associated myelopathy/tropical spastic paraparesis (HAM/TSP) is produced by infection with the retrovirus HTLV-1 (Human T-cell lymphotropic virus type 1 (HTLV-1) [128, 129]. HTLV-1 is transmitted by breast-feeding, sexual intercourse, and parenterally [130]. Worldwide, around 15–20 million people are infected with HTLV-1; however, only 3–5% develop HAM/TSP. Another ~5% develop adult T-cell leukemia/lymphoma (ATL), whereas over 90% of infected people are asymptomatic carriers [131]. The most common HAM/TSP symptom is lower limb motor dysfunction, followed with bladder/bowel dysfunctions and sensory alterations [132]. The virus mainly infects CD4+-T-cells, while monocytes, B-cells, CD8+-T-cells, and DC are infected to a lesser extent and found in spinal cord lesions together with infected astrocytes and endothelial cells [7, 133]. HAM/TSP causes alteration of CNS axonal transport based on the presence of APP deposits in the axons, a classical marker of defects in fast axonal transport [134, 135]. Immunological studies have shown a chronic infiltration of activated CD4+ and CD8+-T-cells into the CNS [136].

It is a consensus that the paraparesis axonopathy generates as a consequence of chronic extracellular action of viral proteins secreted by the infected lymphocytes present in the CNS [137, 138]. Among secreted proteins, Tax viral protein acts on several viral and cellular processes, modulates various cellular signaling pathways, and has also been detected in the CSF of HAM/TSP patients. In the cytoplasm of infected lymphocytes, Tax activates NF-kB pathway responsible for proliferation and differentiation of T-cells, whereas in the nucleus, Tax activates the ATP/CREB pathway. Tax can also be secreted via endoplasmic reticulum-Golgi apparatus and by exosomes [136, 137, 138, 139, 140, 141]. Tax secreted from activated peripheral blood mononuclear cells (PBMCs) could explain the presence of Tax in the plasma and CSF of infected patients and carriers [141].

Using in vitro culture of PBMCs from HAM/TSP patients, it has been recently found that the levels of secreted Sema4D were increased compared to healthy subjects [142]. Elevated Sema4D could be explained as a result of increased levels of MT1-MMP—the enzyme responsible for generating soluble Sema4D (Sema4D) from the transmembrane Sema4D—found in PBMCs from HAM/TSP patients. It has been also found that Tax and SEMA4D co-immunoprecipitate from PBMC culture medium. To test the effect of Tax and Sema4D (or the Tax/Sema4D complex) in neuronal cells, culture media from infected lymphocytes were added to PC12 cells during their differentiation to neuronal type, finding decreased neurite length as a result. The effect of HTLV-1-infected PBMC culture media was blocked by both anti-Sema4D and anti-Tax antibodies, suggesting neurite length reduction by a Tax/Sema4D complex [142]. In the same report, it was shown that infected lymphocytes strongly migrate in response to Sema4D using a trans-well system. It was found that in the population of migrated lymphocytes, the levels of CRMP-2 phosphorylation at Ser522 were increased [142]. A change in Sema4D-mediated phosphorylation of CRMP-2 could be responsible for the increased motility. Authors proposed that infected lymphocytes have an increased migratory response toward Sema4D, making them able to cross the BBB [142]. Once in the CNS, infected lymphocytes secrete Tax and Sema4D, attracting more HTLV-1-infected lymphocytes at the same time that these proteins could mediate pathological disturbances on neuronal cells.

4.6 Multiple sclerosis

MS is a CNS disease mostly considered of autoimmune etiology. It shows demyelinated plaques that sometimes remyelinate spontaneously. Remyelination involves the recruitment of OPC, which differentiate into mature oligodendrocytes in damaged areas to promote remyelination. Nevertheless, the remyelination process is prone to fail, leading to progressive disability [41, 143]. Even though there are multiple reports linking semaphorins with lymphocyte signaling during MS; in this section, we will focus on discussing the reports that have linked semaphorin signaling in oligodendrocytes during MS.

Sema3 proteins are the main semaphorins related to MS, although there is an increasing evidence of Sema4 involvement as well. Using postmortem human samples, the Lubetzki lab [144] showed the presence of numerous cells positive for Sema3A or Sema3F transcripts around and within demyelinating white matter lesions in MS brains, whereas these transcripts were absent in control adult brain white matter. The differential expression of Sema3A and Sema3F was strictly restricted to active plaques. No expression was detected in normal white matter distant to active lesions, around/within chronically demyelinated lesions or remyelinated plaques [144]. Later, also in human MS tissue samples, it was shown that although the chemoattractant Sema3F and chemorepellent Sema3A had similar protein expression patterns in some lesions, Sema3A was predominantly expressed in chronic active lesions, which mostly contain few OPCs [145].

The Lubetzki lab [42] also used a mouse model where demyelinated lesions are induced by lysophosphatidylcholine injection. They found that adult OPCs express Sema3 receptors (plexin A and neuropilin 1 and 2) and that the expression of these receptors, together with Sema3A and Sema3F, is increased after the induction of lesions. Interestingly, in vivo lentiviral expression of Sema3A decreased the OPC density in induced lesions, whereas Sema3F produced the opposite effect. When a transgenic mouse with a mutated NRP1 preventing Sema3A binding was used, an increase in OPC density was found after the induction of lesions compared to wild-type mice. The density of remyelinated axons increased in lesions of animals receiving the Sema3F, but not the Sema3A lentiviral vector [42]. Using a similar approach, in a more recent publication 145, the authors injected recombinant Sema3A and Sema3F to mice. Sema3A-treated mice had significantly fewer OPCs on the side of the lesion compared to the opposite side without lesion, whereas Sema3F-treated mice had increased number of OPCs in the lesion side [145]. Parallel studies in rats have shown that Sema3A inhibits CNS remyelination and the lineage progression of OPCs in demyelinated lesions, arresting OPCs at a premyelinating state [44]. Finally, a recent report using exome sequencing analysis found an association of a missense mutation in the plexin A3 gene (receptor of Sema3A and Sema3F) with increased disability in MS males. Given the gender association, the authors debated whether the plexin A3 mutation could alter the protein stability, interfering with its ligand binding and arguing the possibility of protective effects of estradiol in females [146]. Considering that in MS lesions, Sema3A and its receptors are also expressed in neurons, reactive astrocytes, and microglia/macrophages [147], the source of Sema3A can be multiple and simultaneously affect not only OPCs signaling, but also other cell types.

There are also some reports linking MS with Sema4. Ferritin uptake by oligodendrocytes is mediated by the Tim2 receptor and required for iron acquisition. In addition to ferritin, Tim2 binds Sema4A [148]. Recombinant Sema4A exposure to primary rat OPCs resulted in dose-dependent OPC cytotoxicity. Astrocytes and mature oligodendrocytes were, however, unaffected. The authors suggested that the observed cytotoxicity could be mediated by Tim2 receptor. Lymphocytes, macrophages, or microglia could be the source of Sema4A in vivo [149]. Later, the same group found that human oligodendrocytes undergo apoptosis when exposed to Sema4A and that the levels of this protein are increased in multiple sclerosis patients [150]. A different research group used recombinant Sema4D in an in vitro model of cultured OPC, resulting in actin filament rearrangement indicative of cytoskeletal collapse, along with an increase in apoptotic cells and fewer OPC differentiating into mature oligodendrocytes. All these effects were avoided by incubation with anti-Sema4D antibody [39]. The relative contribution of different semaphorins remains to be tested in future experiments in order to understand their role in the nervous system during MS.

4.7 Cross talk between the immune and the nervous systems

Even though the semaphorin signaling in lymphocytes is not the main subject of this chapter, it is impossible to completely dissociate the semaphorin signaling in the nervous system from their roles in the immune system. The cross talk between these systems is extensive, and different neurological disorders are considered to have an important neuroinflammatory component. For example, the most commonly used animal model for MS is the experimental autoimmune encephalomyelitis (EAE) model, which resembles neuroinflammatory conditions. Different authors suggest that the effect of semaphorins in the nervous system is most likely an indirect effect through modulation of the neuroinflammation produced by immune cells in the nervous system, supported by using the EAE model. For instance, the EAE pathology is exacerbated in Sema7A-deficient mice, and T cells are hyperactive in response to activation in this model. Similarly, Sema4A is increased in MS patients and associated to nonresponsiveness to IFN-β therapy. Anti-Sema4D antibodies inhibited neuroinflammation during EAE [151, 152, 153].

Most of the patients who later will develop MS, usually, have an acute episode of neurological disturbance known as clinically isolated syndrome (CIS). The Montalban lab [154], using mass-spectrometry analysis, identified proteins associated with conversion to MS in CSF samples from CIS patients in a follow-up study. They found that Sema7A was downregulated in patients who later converted to MS [154, 155]. Using the EAE model, the same group found that Sema3A is increased in the CNS and decreased in the immune system, whereas Sema7A is increased in both systems [156]. The above results suggest an intricated system where different semaphorins can be participating at the same time. It is important to understand the relative contribution of different neuronal types and different immune cell types to the pathology and also the amount of soluble semaphorins available to interact with these cells.

Advertisement

5. Conclusion and future perspectives

Throughout this chapter, we have reviewed the currently known implications of different semaphorin classes to some relevant neurological disorders, highlighting their receptors and signaling pathways that could be affected in neuropathologies. Even though the diseases we discussed here represent just a fraction among several other semaphorin-affected neurodegenerative, psychiatric, and immunological disorders, they are also likely representative of the semaphorin function. The advances so far in this field are promising, yet the results obtained from murine systems require testing on human models and subsequently, approaching to eventual therapies and clinical trials. We have already mentioned the potential of semaphorins for cell replacement therapy, such as in the recent approaches on PD [105], or the alternative new drug developments to target specific semaphorin-induced kinases, such as in AD [67]. All these new treatment alternatives emerged in the recent years from the advances in understanding semaphorin-mediated mechanisms on human diseases. In addition, semaphorins have been recently pointed as the center for new therapeutic strategies using blocking antibodies. For example, the VX15/2503, an anti-Sema4D antibody has been characterized for clinical development on MS, Huntington’s disease, and cancer [157]. LaGanke et al. carried out a phase 1 study providing evidence that the VX15/2503 anti-semaphorin 4D antibody administered at various doses was safe and tolerated in patients with MS [158]. It is expected in following years that new breakthroughs will further highlight semaphorin function in neurodegenerative conditions and contribute to future therapeutic strategies.

References

  1. 1. Alto LT, Terman JR. Semaphorins and their signaling mechanisms. Methods in Molecular Biology. 2017;1493:1-25. DOI: 10.1007/978-1-4939-6448-2 1
  2. 2. Pasterkamp RJ, Giger RJ. Semaphorin function in neural plasticity and disease. Current Opinion in Neurobiology. 2009;19:263-274. DOI: 10.1016/j.conb.2009.06.001
  3. 3. Okuno T, Nakatsuji Y, Kumanogoh A. The role of immune semaphorins in multiple sclerosis. FEBS Letters. 2011;585:3829-3835. DOI: 10.1016/j.febslet.2011.03.033
  4. 4. Mecollari V, Nieuwenhuis B, Verhaagen J. A perspective on the role of Class III semaphorin signaling in central nervous system trauma. Frontiers in Cellular Neuroscience. 2014;8:328. DOI: 10.3389/fncel.2014.00328
  5. 5. Van Battum EY, Brignani S, Pasterkamp RJ. Axon guidance proteins in neurological disorders. Lancet Neurology. 2015;14:532-546. DOI: 10.1016/S1474-4422(14)70257-1
  6. 6. Goshima Y, Yamashita N, Nakamura F, Sasaki Y. Regulation of dendritic development by semaphorin 3A through novel intracellular remote signaling. Cell Adhesion & Migration. 2016;10:627-640. DOI: 10.1080/19336918.2016.1210758
  7. 7. Irish BP, Khan ZK, Jain P, Nonnemacher MR, Pirrone V, Rahman S, et al. Molecular mechanisms of neurodegenerative diseases induced by human retroviruses: A review. American Journal of Infectious Diseases. 2009;5:231-258
  8. 8. Gitler AD, Dhillon P, Shorter J. Neurodegenerative disease: Models, mechanisms, and a new hope. Disease Models & Mechanisms. 2017;10:499-502. DOI: 10.1242/dmm.030205
  9. 9. Yazdani U, Terman JR. The semaphorins. Genome Biology. 2006;7:211. DOI: 10.1186/gb-2006-7-3-211
  10. 10. Roney K, Holl E, Ting J. Immune plexins and semaphorins: Old proteins, new immune functions. Protein & Cell. 2013;4:17-26. DOI: 10.1007/s13238-012-2108-4
  11. 11. Yoshida Y. Semaphorin signaling in vertebrate neural circuit assembly. Frontiers in Molecular Neuroscience. 2012;5:71. DOI: 10.3389/fnmol.2012.00071
  12. 12. Zhou Y, Gunput RA, Pasterkamp RJ. Semaphorin signaling: Progress made and promises ahead. Trends in Biochemical Sciences. 2008;33:161-170. DOI: 10.1016/j.tibs.2008.01.006
  13. 13. Sharma A, Verhaagen J, Harvey AR. Receptor complexes for each of the Class 3 Semaphorins. Frontiers in Cellular Neuroscience. 2012;6:28. DOI: 10.3389/fncel.2012.00028. eCollection 2012
  14. 14. Takamatsu H, Kumanogoh A. Diverse roles for semaphorin-plexin signaling in the immune system. Trends in Immunology. 2012;33:127-135. DOI: 10.1016/j.it.2012.01.008
  15. 15. Pasterkamp RJ. Getting neural circuits into shape with semaphorins. Nature Reviews. Neuroscience. 2012;13:605-618. DOI: 10.1038/nrn3302
  16. 16. Tillo M, Ruhrberg C, Mackenzie F. Emerging roles for semaphorins and VEGFs in synaptogenesis and synaptic plasticity. Cell Adhesion & Migration. 2012;6:541-546. DOI: 10.4161/cam.22408
  17. 17. Sarabipour S, Mac Gabhann F. VEGF-A121a binding to Neuropilins-A concept revisited. Cell Adhesion & Migration. 2018;12:204-214. DOI: 10.1080/19336918.2017.1372878
  18. 18. Kimura T, Watanabe H, Iwamatsu A, Kaibuchi K. Tubulin and CRMP-2 complex is transported via Kinesin-1. Journal of Neurochemistry. 2005;93:1371-1382. DOI: 10.1111/j.1471-4159.2005.03063.x
  19. 19. Schmidt EF, Strittmatter SM. The CRMP family of proteins and their role in Sema3A signaling. Advances in Experimental Medicine and Biology. 2007;600:1-11. DOI: 10.1007/978-0-387-70956-7_1
  20. 20. Chae YC, Lee S, Heo K, Ha SH, Jung Y, Kim JH, et al. Collapsin response mediator protein-2 regulates neurite formation by modulating tubulin GTPase activity. Cellular Signalling. 2009;21:1818-1826. DOI: 10.1016/j.cellsig.2009.07.017
  21. 21. Hensley K, Venkova K, Christov A, Gunning W, Park J. Collapsin response mediator protein-2: An emerging pathologic feature and therapeutic target for neurodisease indications. Molecular Neurobiology. 2011;43:180-191. DOI: 10.1007/s12035-011-8166-4
  22. 22. Khanna R, Wilson SM, Brittain JM, Weimer J, Sultana R, Butterfield A, et al. Opening Pandora's jar: A primer on the putative roles of CRMP2 in a panoply of neurodegenerative, sensory and motor neuron, and central disorders. Future Neurology. 2012;7:749-771. DOI: 10.2217/FNL.12.68
  23. 23. Brown JA, Bridgman PC. Disruption of the cytoskeleton during Semaphorin 3A induced growth cone collapse correlates with differences in actin organization and associated binding proteins. Developmental Neurobiology. 2009;69:633-646. DOI: 10.1002/dneu.20732
  24. 24. Oinuma I, Ito Y, Katoh H, Negishi M. Semaphorin 4D/Plexin-B1 stimulates PTEN activity through R-Ras GTPase-activating protein activity, inducing growth cone collapse in hippocampal neurons. The Journal of Biological Chemistry. 2010;285:28200-28209. DOI: 10.1074/jbc.M110.147546
  25. 25. Gallo G. RhoA-kinase coordinates F-actin organization and myosin II activity during semaphorin-3A-induced axon retraction. Journal of Cell Science. 2006;119:3413-3423. DOI: 10.1242/jcs.03084
  26. 26. Dent EW, Gupton SL, Gertler FB. The growth cone cytoskeleton in axon outgrowth and guidance. Cold Spring Harbor Perspectives in Biology. 2011;3:1-39. pii: a001800. DOI: 10.1101/cshperspect.a001800
  27. 27. Vodrazka P, Korostylev A, Hirschberg A, Swiercz JM, Worzfeld T, Deng S, et al. The semaphorin 4D-plexin-B signalling complex regulates dendritic and axonal complexity in developing neurons via diverse pathways. The European Journal of Neuroscience. 2009;30:1193-1208. DOI: 10.1111/j.1460-9568.2009.06934.x
  28. 28. McDermott JE, Goldblatt D, Paradis S. Class 4 Semaphorins and Plexin-B receptors regulate GABAergic and glutamatergic synapse development in the mammalian hippocampus. Molecular and Cellular Neurosciences. 2018;92:50-66. DOI: 10.1016/j.mcn.2018.06.008
  29. 29. Afshari FT, Kappagantula S, Fawcett JW. Extrinsic and intrinsic factors controlling axonal regeneration after spinal cord injury. Expert Reviews in Molecular Medicine. 2009;11:e37. DOI: 10.1017/S1462399409001288
  30. 30. Sandvig A, Berry M, Barrett LB, Butt A, Logan A. Myelin-, reactive glia-, and scar-derived CNS axon growth inhibitors: Expression, receptor signaling, and correlation with axon regeneration. Glia. 2004;46:225-251. DOI: 10.1002/glia.10315
  31. 31. Bolsover S, Fabes J, Anderson PN. Axonal guidance molecules and the failure of axonal regeneration in the adult mammalian spinal cord. Restorative Neurology and Neuroscience. 2008;26:117-130
  32. 32. McCormick AM, Leipzig ND. Neural regenerative strategies incorporating biomolecular axon guidance signals. Annals of Biomedical Engineering. 2012;40:578-597. DOI: 10.1007/s10439-011-0505-0
  33. 33. Shim SO, Cafferty WB, Schmidt EC, Kim BG, Fujisawa H, Strittmatter SM. PlexinA2 limits recovery from corticospinal axotomy by mediating oligodendrocyte-derived Sema6A growth inhibition. Molecular and Cellular Neurosciences. 2012;50:193-200. DOI: 10.1016/j.mcn.2012.04.007
  34. 34. Zhang M, Zhou Q, Luo Y, Nguyen T, Rosenblatt MI, Guaiquil VH. Semaphorin3A induces nerve regeneration in the adult cornea-a switch from its repulsive role in development. PLoS One. 2018;13:e0191962. DOI: 10.1371/journal.pone.0191962. eCollection 2018
  35. 35. Anderson JE, Do MQ, Daneshvar N, Suzuki T, Dort J, Mizunoya W, et al. The role of semaphorin3A in myogenic regeneration and the formation of functional neuromuscular junctions on new fibres. Biological Reviews of the Cambridge Philosophical Society. 2017;92:1389-1405. DOI: 10.1111/brv.12286
  36. 36. Peng SX, Yao L, Cui C, Zhao HD, Liu CJ, Li YH, et al. Semaphorin4D promotes axon regrowth and swimming ability during recovery following zebrafish spinal cord injury. Neuroscience. 2017;351:36-46. DOI: 10.1016/j.neuroscience.2017.03.030
  37. 37. Sakurai A, Doçi CL, Gutkind JS. Semaphorin signaling in angiogenesis, lymphangiogenesis and cancer. Cell Research. 2012;22:23-32. DOI: 10.1038/cr.2011.198
  38. 38. Oh WJ, Gu C. The role and mechanism-of-action of Sema3E and Plexin-D1 in vascular and neural development. Seminars in Cell & Developmental Biology. 2013;24:156-162. DOI: 10.1016/j.semcdb.2012.12.001
  39. 39. Smith ES, Jonason A, Reilly C, Veeraraghavan J, Fisher T, Doherty M, et al. SEMA4D compromises blood-brain barrier, activates microglia, and inhibits remyelination in neurodegenerative disease. Neurobiology of Disease. 2015;73:254-268. DOI: 10.1016/j.nbd.2014.10.008
  40. 40. Basile JR, Gavard J, Gutkind JS. Plexin-B1 utilizes RhoA and Rho kinase to promote the integrin-dependent activation of Akt and ERK and endothelial cell motility. The Journal of Biological Chemistry. 2007;282:34888-34895. DOI: 10.1074/jbc.M705467200
  41. 41. Franklin RJ, Ffrench-Constant C. Remyelination in the CNS: From biology to therapy. Nature Reviews. Neuroscience. 2008;9:839-855. DOI: 10.1038/nrn2480
  42. 42. Piaton G, Aigrot MS, Williams A, Moyon S, Tepavcevic V, Moutkine I, et al. Class 3 semaphorins influence oligodendrocyte precursor recruitment and remyelination in adult central nervous system. Brain. 2011;134:1156-1167. DOI: 10.1093/brain/awr022
  43. 43. Alizadeh A, Dyck SM, Karimi-Abdolrezaee S. Myelin damage and repair in pathologic CNS: Challenges and prospects. Frontiers in Molecular Neuroscience. 2015;8:35. DOI: 10.3389/fnmol.2015.00035
  44. 44. Syed YA, Hand E, Möbius W, Zhao C, Hofer M, Nave KA, et al. Inhibition of CNS remyelination by the presence of semaphorin 3A. The Journal of Neuroscience. 2011;31:3719-3728. DOI: 10.1523/JNEUROSCI.4930-10.2011
  45. 45. Yamaguchi W, Tamai R, Kageura M, Furuyama T, Inagaki S. Sema4D as an inhibitory regulator in oligodendrocyte development. Molecular and Cellular Neurosciences. 2012;49:290-299. DOI: 10.1016/j.mcn.2011.12.004
  46. 46. Takegahara N, Kumanogoh A, Kikutani H. Semaphorins: A new class of immunoregulatory molecules. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences. 2005;360:1673-1680. DOI: 10.1098/rstb.2005.1696
  47. 47. Nkyimbeng-Takwi E, Chapoval SP. Biology and function of neuroimmune semaphorins 4A and 4D. Immunologic Research. 2011;50:10-21. DOI: 10.1007/s12026-010-8201-y
  48. 48. Tian L, Rauvala H, Gahmberg CG. Neuronal regulation of immune responses in the central nervous system. Trends in Immunology. 2009;30:91-99. DOI: 10.1016/j.it.2008.11.002
  49. 49. Chavarría A, Cárdenas G. Neuronal influence behind the central nervous system regulation of the immune cells. Frontiers in Integrative Neuroscience. 2013;7:64. DOI: 10.3389/fnint.2013.00064
  50. 50. Lepelletier Y, Moura IC, Hadj-Slimane R, Renand A, Fiorentino S, Baude C, et al. Immunosuppressive role of semaphorin-3A on T cell proliferation is mediated by inhibition of actin cytoskeleton reorganization. European Journal of Immunology. 2006;36:1782-1793. DOI: 10.1002/eji.200535601
  51. 51. Pellet-Many C, Frankel P, Jia H, Zachary I. Neuropilins: Structure, function and role in disease. Biochemical Journal. 2008;411:211-226. DOI: 10.1042/BJ20071639
  52. 52. Rezaeepoor M, Shapoori S, Ganjalikhani-Hakemi M, Etemadifar M, Alsahebfosoul F, Eskandari N, et al. Decreased expression of Sema3A, an immune modulator, in blood sample of multiple sclerosis patients. Gene. 2017;610:59-63. DOI: 10.1016/j.gene.2017.02.013
  53. 53. Burns A, Iliffe S. Alzheimer’s disease. British Medical Journal. 2009;338:b158. DOI: 10.1136/bmj.b158
  54. 54. Giger RJ, Pasterkamp RJ, Heijnen S, Holtmaat AJ, Verhaagen J. Anatomical distribution of the chemorepellent semaphorin III/collapsin-1 in the adult rat and human brain: Predominant expression in structures of the olfactory-hippocampal pathway and the motor system. Journal of Neuroscience Research. 1998;52:27-42. DOI: 10.1002/(SICI)1097-4547(19980401)52:1<27::AID-JNR4>3.0.CO;2-M
  55. 55. Hirsch E, Hu LJ, Prigent A, Constantin B, Agid Y, Drabkin H, et al. Distribution of semaphorin IV in adult human brain. Brain Research. 1999;823:67-79
  56. 56. Santa-Maria I, Avila J, Rabano A. Differential gene expression analysis of human entorhinal cortex support a possible role of some extracellular matrix proteins in the onset of Alzheimer disease. Neuroscience Letters. 2010;468:225-228. DOI: 10.1016/j.neulet.2009.11.002
  57. 57. Good PF, Alapat D, Hsu A, Chu C, Perl D, Wen X, et al. A role for semaphorin 3A signaling in the degeneration of hippocampal neurons during Alzheimer’s disease. Journal of Neurochemistry. 2004;91:716-736. DOI: 10.1111/j.1471-4159.2004.02766.x
  58. 58. Yoshida H, Watanabe A, Ihara Y. Collapsin response mediator protein-2 is associated with neurofibrillary tangles in Alzheimer’s disease. The Journal of Biological Chemistry. 1998;273:9761-9768
  59. 59. Goshima Y, Nakamura F, Strittmatter P, Strittmatter SM. Collapsin-induced growth cone collapse mediated by an intracellular protein related to UNC-33. Nature. 1995;376:509-514. DOI: 10.1038/376509a0
  60. 60. Gu Y, Hamajima N, Ihara Y. Neurofibrillary tangle-associated collapsin response mediator protein-2 (CRMP-2) is highly phosphorylated on Thr-509, Ser-518, and Ser-522. Biochemistry. 2000;39:4267-4275
  61. 61. Cole AR, Knebel A, Morrice NA, Robertson LA, Irving AJ, Connolly CN, et al. GSK-3 phosphorylation of the Alzheimer epitope within collapsin response mediator proteins regulates axon elongation in primary neurons. The Journal of Biological Chemistry. 2004;279:50176-50180. DOI: 10.1074/jbc.C400412200
  62. 62. Jun G, Asai H, Zeldich E, Drapeau E, Chen C, Chung J, et al. PLXNA4 is associated with Alzheimer disease and modulates tau phosphorylation. Annals of Neurology. 2014;76:379-392. DOI: 10.1002/ana.24219
  63. 63. Brown M, Jacobs T, Eickholt B, Ferrari G, Teo M, Monfries C, et al. Alpha2-chimaerin, cyclin-dependent Kinase 5/p35, and its target collapsin response mediator protein-2 are essential components in semaphorin 3A-induced growth-cone collapse. The Journal of Neuroscience. 2004;24:8994-9004. DOI: 10.1523/JNEUROSCI.3184-04.2004
  64. 64. Uchida Y, Ohshima T, Sasaki Y, Suzuki H, Yanai S, Yamashita N, et al. Semaphorin3A signalling is mediated via sequential Cdk5 and GSK3beta phosphorylation of CRMP2: Implication of common phosphorylating mechanism underlying axon guidance and Alzheimer’s disease. Genes to Cells. 2005;10:165-179. DOI: 10.1111/j.1365-2443.2005.00827.x
  65. 65. Sasaki Y, Cheng C, Uchida Y, Nakajima O, Ohshima T, Yagi T, et al. Fyn and Cdk5 mediate semaphorin-3A signaling, which is involved in regulation of dendrite orientation in cerebral cortex. Neuron. 2002;35:907-920
  66. 66. Noble W, Olm V, Takata K, Casey E, Mary O, Meyerson J, et al. Cdk5 is a key factor in tau aggregation and tangle formation in vivo. Neuron. 2003;38:555-565
  67. 67. Nygaard HB, Wagner AF, Bowen GS, Good SP, MacAvoy MG, Strittmatter KA, et al. A phase Ib multiple ascending dose study of the safety, tolerability, and central nervous system availability of AZD0530 (saracatinib) in Alzheimer’s disease. Alzheimer's Research & Therapy. 2015;7:35. DOI: 10.1186/s13195-015-0119-0
  68. 68. Romberg C, Yang S, Melani R, Andrews MR, Horner AE, Spillantini MG, et al. Depletion of perineuronal nets enhances recognition memory and long-term depression in the perirhinal cortex. The Journal of Neuroscience. 2013;33:7057-7065. DOI: 10.1523/JNEUROSCI.6267-11.2013
  69. 69. Yang S, Hilton S, Alves JN, Saksida LM, Bussey T, Matthews RT, et al. Antibody recognizing 4-sulfated chondroitin sulfate proteoglycans restores memory in tauopathy-induced neurodegeneration. Neurobiology of Aging. 2017;59:197-209. DOI: 10.1016/j.neurobiolaging.2017.08.002
  70. 70. Dick G, Tan CL, Alves JN, Ehlert EME, Miller GM, Hsieh-Wilson LC, et al. Semaphorin 3A binds to the perineuronal nets via chondroitin sulfate type E motifs in rodent brains. The Journal of Biological Chemistry. 2013;288:27384-27395. DOI: 10.1074/jbc.M111.310029
  71. 71. Moreno-Flores MT, Martín-Aparicio E, Martín-Bermejo MJ, Agudo M, McMahon S, Avila J, et al. Semaphorin 3C preserves survival and induces neuritogenesis of cerebellar granule neurons in culture. Journal of Neurochemistry. 2003;87:879-890
  72. 72. Hauser PS, Ryan RO. Impact of apolipoprotein E on Alzheimer’s disease. Current Alzheimer Research. 2013;10:809-817
  73. 73. Villa C, Venturelli E, Fenoglio C, De Riz M, Scalabrini D, Cortini F, et al. Candidate gene analysis of semaphorins in patients with Alzheimer’s disease. Neurological Sciences. 2010;31:169-173. DOI: 10.1007/s10072-009-0200-1
  74. 74. Ueki M, Kawasaki Y, Tamiya G. For Alzheimer’s disease neuroimaging initiative. Detecting genetic association through shortest paths in a bidirected graph. Genetic Epidemiology. 2017;41:481-497. DOI: 10.1002/gepi.22051
  75. 75. Schott JM, Crutch SJ, Carrasquillo MM, Uphill J, Shakespeare TJ, Ryan NS, et al. Genetic risk factors for the posterior cortical atrophy variant of Alzheimer’s disease. Alzheimer's & Dementia. 2016;12:862-871. DOI: 10.1016/j.jalz.2016.01.010
  76. 76. Tysnes O-B, Storstein A. Epidemiology of Parkinson’s disease. Journal of Neural Transmission (Vienna). 2017;124:901-905. DOI: 10.1007/s00702-017-1686-y
  77. 77. Shirvan A, Ziv I, Fleminger G, Shina R, He Z, Brudo I, et al. Semaphorins as mediators of neuronal apoptosis. Journal of Neurochemistry. 1999;73:961-971
  78. 78. Shirvan A, Shina R, Ziv I, Melamed E, Barzilai A. Induction of neuronal apoptosis by Semaphorin3A-derived peptide. Brain Research. Molecular Brain Research. 2000;83:81-93
  79. 79. Barzilai A, Zilkha-Falb R, Daily D, Stern N, Offen D, Ziv I, et al. The molecular mechanism of dopamine-induced apoptosis: Identification and characterization of genes that mediate dopamine toxicity. Journal of Neural Transmission. Supplementum. 2000;60:59-76
  80. 80. Barzilai A, Daily D, Zilkha-Falb R, Ziv I, Offen D, Melamed E, et al. The molecular mechanisms of dopamine toxicity. Advances in Neurology. 2003;91:73-82
  81. 81. Rocca WA, McDonnell SK, Strain KJ, Bower JH, Ahlskog JE, Elbaz A, et al. Familial aggregation of Parkinson’s disease: The Mayo clinic family study. Annals of Neurology. 2004;56:495-502. DOI: 10.1002/ana.20228
  82. 82. Maraganore DM, de Andrade M, Lesnick TG, Strain KJ, Farrer MJ, Rocca WA, et al. High-resolution whole-genome association study of Parkinson disease. American Journal of Human Genetics. 2005;77:685-693. DOI: 10.1086/496902
  83. 83. Bialecka M, Kurzawski M, Klodowska-Duda G, Opala G, Tan E-K, Drozdzik M. Polymorphism in semaphorin 5A (Sema5A) gene is not a marker of Parkinson’s disease risk. Neuroscience Letters. 2006;399:121-123. DOI: 10.1016/j.neulet.2006.01.038
  84. 84. Elbaz A, Nelson LM, Payami H, Ioannidis JPA, Fiske BK, Annesi G, et al. Lack of replication of thirteen single-nucleotide polymorphisms implicated in Parkinson’s disease: A large-scale international study. Lancet Neurology. 2006;5:917-923. DOI: 10.1016/S1474-4422(06)70579-8
  85. 85. Clarimon J, Scholz S, Fung H-C, Hardy J, Eerola J, Hellstrom O, et al. Conflicting results regarding the semaphorin gene (SEMA5A) and the risk for Parkinson disease. American Journal of Human Genetics. 2006;78:1082-1084; author reply 1092-1094. DOI: 10.1086/504727
  86. 86. Ding H, Wang F, Ding X, Song X, Lu X, Zhang K, et al. Association study of semaphorin 5A with risk of Parkinson’s disease in a Chinese Han population. Brain Research. 2008;1245:126-129. DOI: 10.1016/j.brainres.2008.09.080
  87. 87. Yu X, Wang F, Zhang J-P. Meta analysis of the association of rs7702187 SNP in SEMA5A gene with risk of Parkinson’s disease. European Review for Medical and Pharmacological Sciences. 2014;18:900-904
  88. 88. Yasuhara T, Shingo T, Kobayashi K, Takeuchi A, Yano A, Muraoka K, et al. Neuroprotective effects of vascular endothelial growth factor (VEGF) upon dopaminergic neurons in a rat model of Parkinson’s disease. The European Journal of Neuroscience. 2004;19:1494-1504. DOI: 10.1111/j.1460-9568.2004.03254.x
  89. 89. Yasuhara T, Shingo T, Muraoka K, Kameda M, Agari T, Wenji Y, et al. Toxicity of semaphorin3A for dopaminergic neurons. Neuroscience Letters. 2005;382:61-65. DOI: 10.1016/j.neulet.2005.02.064
  90. 90. Rodriguez-Perez AI, Dominguez-Meijide A, Lanciego JL, Guerra MJ, Labandeira-Garcia JL. Inhibition of Rho kinase mediates the neuroprotective effects of estrogen in the MPTP model of Parkinson’s disease. Neurobiology of Disease. 2013;58:209-219. DOI: 10.1016/j.nbd.2013.06.004
  91. 91. Zhao Y, Zhang Q, Xi J, Li Y, Ma C, Xiao B. Multitarget intervention of Fasudil in the neuroprotection of dopaminergic neurons in MPTP-mouse model of Parkinson’s disease. Journal of the Neurological Sciences. 2015;353:28-37. DOI: 10.1016/j.jns.2015.03.022
  92. 92. Qi L, Tang Y-G, Wang L, He W, Pan H-H, Nie R-R, et al. Role of Rho-mediated ROCK-Semaphorin3A signaling pathway in the pathogenesis of Parkinson’s disease in a mouse model. Journal of the Neurological Sciences. 2016;370:21-26. DOI: 10.1016/j.jns.2016.08.061
  93. 93. Rodriguez-Pallares J, Rodriguez-Perez AI, Muñoz A, Parga JA, Toledo-Aral JJ, Labandeira-Garcia JL. Effects of Rho kinase inhibitors on grafts of dopaminergic cell precursors in a rat model of Parkinson’s disease. Stem Cells Translational Medicine. 2016;5:804-815. DOI: 10.5966/sctm.2015-0182
  94. 94. Grünblatt E, Mandel S, Youdim MB. Neuroprotective strategies in Parkinson’s disease using the models of 6-hydroxydopamine and MPTP. Annals of the New York Academy of Sciences. 2000;899:262-273
  95. 95. Pacelli C, Giguère N, Bourque M-J, Lévesque M, Slack RS, Trudeau L-É. Elevated mitochondrial bioenergetics and axonal arborization size are key contributors to the vulnerability of dopamine neurons. Current Biology. 2015;25:2349-2360. DOI: 10.1016/j.cub.2015.07.050
  96. 96. Hernández-Montiel HL, Tamariz E, Sandoval-Minero MT, Varela-Echavarría A. Semaphorins 3A, 3C, and 3F in mesencephalic dopaminergic axon pathfinding. The Journal of Comparative Neurology. 2008;506:387-397. DOI: 10.1002/cne.21503
  97. 97. Tamariz E, Díaz-Martínez NE, Díaz NF, García-Peña CM, Velasco I, Varela-Echavarría A. Axon responses of embryonic stem cell-derived dopaminergic neurons to semaphorins 3A and 3C. Journal of Neuroscience Research. 2010;88:971-980. DOI: 10.1002/jnr.22268
  98. 98. Díaz-Martínez NE, Tamariz E, Díaz NF, García-Peña CM, Varela-Echavarría A, Velasco I. Recovery from experimental parkinsonism by semaphorin-guided axonal growth of grafted dopamine neurons. Molecular Therapy. 2013;21:1579-1591. DOI: 10.1038/mt.2013.78
  99. 99. Carballo-Molina OA, Sánchez-Navarro A, López-Ornelas A, Lara-Rodarte R, Salazar P, Campos-Romo A, et al. Semaphorin 3C released from a biocompatible hydrogel guides and promotes axonal growth of rodent and human dopaminergic neurons. Tissue Engineering. Part A. 2016;22:850-861. DOI: 10.1089/ten.TEA.2016.0008
  100. 100. Kim J-H, Auerbach JM, Rodríguez-Gómez JA, Velasco I, Gavin D, Lumelsky N, et al. Dopamine neurons derived from embryonic stem cells function in an animal model of Parkinson’s disease. Nature. 2002;418:50-56. DOI: 10.1038/nature00900
  101. 101. Isacson O, Bjorklund LM, Schumacher JM. Toward full restoration of synaptic and terminal function of the dopaminergic system in Parkinson’s disease by stem cells. Annals of Neurology. 2003;53(Suppl 3):S135-S146; discussion S146-148. DOI: 10.1002/ana.10482
  102. 102. Rodríguez-Gómez JA, Lu JQ, Velasco I, Rivera S, Zoghbi SS, Liow JS, et al. Persistent dopamine functions of neurons derived from embryonic stem cells in a rodent model of Parkinson disease. Stem Cells. 2007;25:918-928. DOI: 10.1634/stemcells.2006-0386
  103. 103. Wernig M, Zhao J-P, Pruszak J, Hedlund E, Fu D, Soldner F, et al. Neurons derived from reprogrammed fibroblasts functionally integrate into the fetal brain and improve symptoms of rats with Parkinson’s disease. Proceedings of the National Academy of Sciences of the United States of America. 2008;105:5856-5861. DOI: 10.1073/pnas.0801677105
  104. 104. García-Peña CM, Kim M, Frade-Pérez D, Avila-González D, Téllez E, Mastick GS, et al. Ascending midbrain dopaminergic axons require descending GAD65 axon fascicles for normal pathfinding. Frontiers in Neuroanatomy. 2014;8:43. DOI: 10.3389/fnana.2014.00043
  105. 105. Chen D, Fu W, Zhuang W, Lv C, Li F, Wang X. Therapeutic effects of intranigral transplantation of mesenchymal stem cells in rat models of Parkinson’s disease. Journal of Neuroscience Research. 2017;95:907-917. DOI: 10.1002/jnr.23879
  106. 106. Roy NS, Cleren C, Singh SK, Yang L, Beal MF, Goldman SA. Functional engraftment of human ES cell-derived dopaminergic neurons enriched by coculture with telomerase-immortalized midbrain astrocytes. Nature Medicine. 2006;12:1259-1268. DOI: 10.1038/nm1495
  107. 107. Saporta AS, Sottile SL, Miller LJ, Feely SM, Siskind CE, Shy ME. Charcot-Marie-Tooth disease subtypes and genetic testing strategies. Annals of Neurology. 2011;69:22-33. DOI: 10.1002/ana.22166
  108. 108. Rossor AM, Polke JM, Houlden H, Reilly MM. Clinical implications of genetic advances in Charcot-Marie-Tooth disease. Nature Reviews. Neurology. 2013;9:562-571. DOI: 10.1038/nrneurol.2013.179
  109. 109. Pisciotta C, Shy ME. Neuropathy. Handbook of Clinical Neurology. 2018;148:653-665. DOI: 10.1016/B978-0-444-64076-5.00042-9
  110. 110. Kazamel M, Boes CJ. Charcot-Marie-Tooth disease (CMT): Historical perspectives and evolution. Journal of Neurology. 2015;262:801-805. DOI: 10.1007/s00415-014-7490-9
  111. 111. Antonellis A, Ellsworth RE, Sambuughin N, Puls I, Abel A, Lee-Lin SQ, et al. Glycyl tRNA synthetase mutations in Charcot-Marie-Tooth disease type 2D and distal spinal muscular atrophy type V. American Journal of Human Genetics. 2003;72:1293-1299. DOI: 10.1086/375039
  112. 112. He W, Bai G, Zhou H, Wei N, White NM, Lauer J, et al. CMT2D neuropathy is linked to the neomorphic binding activity of glycyl-tRNA synthetase. Nature. 2015;526:710-714. DOI: 10.1038/nature15510
  113. 113. Park MC, Kang T, Jin D, Han JM, Kim SB, Park YJ, et al. Secreted human glycyl-tRNA synthetase implicated in defense against ERK-activated tumorigenesis. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:E640-E647. DOI: 10.1073/pnas.1200194109
  114. 114. Grice SJ, Sleigh JN, Motley WW, Liu JL, Burgess RW, Talbot K, et al. Dominant, toxic gain-of-function mutations in gars lead to non-cell autonomous neuropathology. Human Molecular Genetics. 2015;24:4397-4406. DOI: 10.1093/hmg/ddv176
  115. 115. Grice SJ, Sleigh JN, Zameel Cader M. Plexin-Semaphorin signaling modifies neuromuscular defects in a drosophila model of peripheral neuropathy. Frontiers in Molecular Neuroscience. 2018;11:55. DOI: 10.3389/fnmol.2018.00055
  116. 116. Sleigh JN, Gómez-Martín A, Wei N, Bai G, Yang XL, Schiavo G. Neuropilin 1 sequestration by neuropathogenic mutant glycyl-tRNA synthetase is permissive to vascular homeostasis. Scientific Reports. 2017;7:9216. DOI: 10.1038/s41598-017-10005-w
  117. 117. Mo Z, Zhao X, Liu H, Hu Q, Chen XQ, Pham J, et al. Aberrant GlyRS-HDAC6 interaction linked to axonal transport deficits in Charcot-Marie-Tooth neuropathy. Nature Communications. 2018;9:1007. DOI: 10.1038/s41467-018-03461-z
  118. 118. Schmidt ER, Pasterkamp RJ, van den Berg LH. Axon guidance proteins: Novel therapeutic targets for ALS? Progress in Neurobiology. 2009;88:286-301. DOI: 10.1016/j.pneurobio.2009.05.004
  119. 119. De Winter F, Vo T, Stam FJ, Wisman LA, Bär PR, Niclou SP, et al. The expression of the chemorepellent Semaphorin 3A is selectively induced in terminal Schwann cells of a subset of neuromuscular synapses that display limited anatomical plasticity and enhanced vulnerability in motor neuron disease. Molecular and Cellular Neurosciences. 2006;32:102-117. DOI: 10.1016/j.mcn.2006.03.002
  120. 120. Moloney EB, Hobo B, De Winter F, Verhaagen J. Expression of a mutant SEMA3A protein with diminished signalling capacity does not alter ALS-related motor decline, or confer changes in NMJ plasticity after BotoxA-induced paralysis of male gastrocnemic muscle. PLoS One. 2017;12:e0170314. DOI: 10.1371/journal.pone.0170314
  121. 121. Venkova K, Christov A, Kamaluddin Z, Kobalka P, Siddiqui S, Hensley K. Semaphorin 3A signaling through neuropilin-1 is an early trigger for distal axonopathy in the SOD1G93A mouse model of amyotrophic lateral sclerosis. Journal of Neuropathology and Experimental Neurology. 2014;73:702-713. DOI: 10.1097/NEN.0000000000000086
  122. 122. Miyazaki K, Nagai M, Morimoto N, Kurata T, Takehisa Y, Ikeda Y, et al. Spinal anterior horn has the capacity to self-regenerate in amyotrophic lateral sclerosis model mice. Journal of Neuroscience Research. 2009;87:3639-3648. DOI: 10.1002/jnr.22156
  123. 123. Lobsiger CS, Boillée S, Cleveland DW. Toxicity from different SOD1 mutants dysregulates the complement system and the neuronal regenerative response in ALS motor neurons. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:7319-7326. DOI: 10.1073/pnas.0702230104
  124. 124. Haidar M, Rchiad Z, Ansari HR, Ben-Rached F, Tajeri S, Latre De Late P, et al. miR-126-5p by direct targeting of JNK-interacting protein-2 (JIP-2) plays a key role in Theileria-infected macrophage virulence. PLoS Pathogens. 2018;14:e1006942. DOI: 10.1371/journal.ppat.1006942
  125. 125. Rotem N, Magen I, Ionescu A, Gershoni-Emek N, Altman T, Costa CJ, et al. ALS along the axons—Expression of coding and noncoding RNA differs in axons of ALS models. Scientific Reports. 2017;7:44500. DOI: 10.1038/srep44500
  126. 126. Maimon R, Ionescu A, Bonnie A, Sweetat S, Wald-Altman S, Inbar S, et al. miR126-5p downregulation facilitates axon degeneration and NMJ disruption via a non-cell-autonomous mechanism in ALS. The Journal of Neuroscience. 2018;38:5478-5494. DOI: 10.1523/JNEUROSCI.3037-17.2018
  127. 127. Körner S, Böselt S, Wichmann K, Thau-Habermann N, Zapf A, Knippenberg S, et al. The axon guidance protein Semaphorin 3A is increased in the motor cortex of patients with amyotrophic lateral sclerosis. Journal of Neuropathology and Experimental Neurology. 2016;75:326-333. pii: nlw003. DOI: 10.1093/jnen/nlw003
  128. 128. Nagai M, Osame M. Human T-cell lymphotropic virus type I and neurological diseases. Journal of Neurovirology. 2003;9:228-235. DOI: 10.1080/13550280390194028
  129. 129. Gessain A, Cassar O. Epidemiological aspects and world distribution of HTLV-1 infection. Frontiers in Microbiology. 2012;3:388. DOI: 10.3389/fmicb.2012.00388
  130. 130. Bangham CRM, Matsuoka M. Human T-cell leukaemia virus type 1: Parasitism and pathogenesis. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences. 2017;372:1-10. pii: 20160272. DOI: 10.1098/rstb.2016.0272
  131. 131. Edlich R, Hill LG, Williams FM. An update on the global epidemic of human T-cell lymphotrophic virus type-I (HTLV-I). Journal of Long-Term Effects of Medical Implants. 2017;27:355-368. DOI: 10.1615/JLongTermEffMedImplants.v27.i2-4.160
  132. 132. Sato T, Yagishita N, Tamaki K, Inoue E, Hasegawa D, Nagasaka M, et al. Proposal of classification criteria for HTLV-1-associated myelopathy/tropical spastic paraparesis disease activity. Frontiers in Microbiology. 2018;9:1651. DOI: 10.3389/fmicb.2018.01651
  133. 133. Echevarria-Lima J, de Abreu Pereira D, de Oliveira TS, de Melo Espíndola O, Lima MA, Celestino Leite AC, et al. Protein profile of blood monocytes is altered in HTLV-1 infected patients: Implications for HAM/TSP disease. Scientific Reports. 2018;8:14354. DOI: 10.1038/s41598-018-32324-2
  134. 134. Coleman M. Axon degeneration mechanisms: Commonality amid diversity. Nature Reviews. Neuroscience. 2005;6:889-898. DOI: 10.1038/nrn1788
  135. 135. Cartier L, Vergara C, Valenzuela MA. Immunohistochemistry of degenerative changes in the central nervous system in spastic paraparesis associated to human T lymphotropic virus type I (HTLV-I). Revista Médica de Chile. 2007;135:1139-1146. DOI: S0034-98872007000900007
  136. 136. Anderson MR, Pleet ML, Enose-Akahata Y, Erickson J, Monaco MC, Akpamagbo Y, et al. Viral antigens detectable in CSF exosomes from patients with retrovirus associated neurologic disease: Functional role of exosomes. Clinical and Translational Medicine. 2018;7:24. DOI: 10.1186/s40169-018-0204-7
  137. 137. Cartier L, Ramirez E. Presence of HTLV-I tax protein in cerebrospinal fluid from HAM/TSP patients. Archives of Virology. 2005;150:743-753. DOI: 10.1007/s00705-004-0443-3
  138. 138. Jain P, Ahuja J, Khan ZK, Shimizu S, Meucci O, Jennings SR, et al. Modulation of dendritic cell maturation and function by the tax protein of human T cell leukemia virus type 1. Journal of Leukocyte Biology. 2007;82:44-56. DOI: 10.1189/jlb.1006641
  139. 139. Alefantis T, Jain P, Ahuja J, Mostoller K, Wigdahl B. HTLV-1 tax nucleocytoplasmic shuttling, interaction with the secretory pathway, extracellular signaling, and implications for neurologic disease. Journal of Biomedical Science. 2005;12:961-974. DOI: 10.1007/s11373-005-9026-x
  140. 140. Medina F, Quintremil S, Alberti C, Barriga A, Cartier L, Puente J, et al. Tax posttranslational modifications and interaction with calreticulin in MT-2 cells and human peripheral blood mononuclear cells of human T cell lymphotropic virus type-I-associated myelopathy/tropical spastic paraparesis patients. AIDS Research and Human Retroviruses. 2014;30:370-379. DOI: 10.1089/AID.2013.0036
  141. 141. Medina F, Quintremil S, Alberti C, Godoy F, Pando ME, Bustamante A, et al. Tax secretion from peripheral blood mononuclear cells and tax detection in plasma of patients with human T-lymphotropic virus-type 1-associated myelopathy/tropical spastic paraparesis and asymptomatic carriers. Journal of Medical Virology. 2016;88:521-531. DOI: 10.1002/jmv.24342
  142. 142. Quintremil S, Alberti C, Rivera M, Medina F, Puente J, Cartier L, et al. Tax and Semaphorin 4D released from lymphocytes infected with human lymphotropic virus type 1 and their effect on neurite growth. AIDS Research and Human Retroviruses. 2016;32:68-79. DOI: 10.1089/aid.2015.0008
  143. 143. Patani R, Balaratnam M, Vora A, Reynolds R. Remyelination can be extensive in multiple sclerosis despite a long disease course. Neuropathology and Applied Neurobiology. 2007;33:277-287. DOI: 10.1111/j.1365-2990.2007.00805.x
  144. 144. Williams A, Piaton G, Aigrot MS, Belhadi A, Théaudin M, Petermann F, et al. Semaphorin 3A and 3F: Key players in myelin repair in multiple sclerosis? Brain. 2007;130:2554-2565. DOI: 10.1093/brain/awm202
  145. 145. Boyd A, Zhang H, Williams A. Insufficient OPC migration into demyelinated lesions is a cause of poor remyelination in MS and mouse models. Acta Neuropathologica. 2013;125:841-859. DOI: 10.1007/s00401-013-1112-y
  146. 146. Qureshi M, Hatem M, Alroughani R, Jacob SP, Al-Temaimi RA. PLXNA3 variant rs5945430 is associated with severe clinical course in male multiple sclerosis patients. Neuromolecular Medicine. 2017;19:286-292. DOI: 10.1007/s12017-017-8443-0
  147. 147. Costa C, Martínez-Sáez E, Gutiérrez-Franco A, Eixarch H, Castro Z, Ortega-Aznar A, et al. Expression of semaphorin 3A, semaphorin 7A and their receptors in multiple sclerosis lesions. Multiple Sclerosis. 2015;21:1632-1643. DOI: 10.1177/1352458515599848
  148. 148. Kumanogoh A, Marukawa S, Suzuki K, Takegahara N, Watanabe C, Ch'ng E, et al. Class IV semaphorin Sema4A enhances T-cell activation and interacts with Tim-2. Nature. 2002;419:629-633. DOI: 10.1038/nature01037
  149. 149. Leitner DF, Todorich B, Zhang X, Connor JR. Semaphorin4A is cytotoxic to oligodendrocytes and is elevated in microglia and multiple sclerosis. ASN Neuro. 2015;7:1-13. pii: 1759091415587502. DOI: 10.1177/1759091415587502
  150. 150. Chiou B, Lucassen E, Sather M, Kallianpur A, Connor J. Semaphorin4A and H-ferritin utilize Tim-1 on human oligodendrocytes: A novel neuro-immune axis. Glia. 2018;66:1317-1330. DOI: 10.1002/glia.23313
  151. 151. Czopik AK, Bynoe MS, Palm N, Raine CS, Medzhitov R. Semaphorin 7A is a negative regulator of T cell responses. Immunity. 2006;24:591-600. DOI: 10.1016/j.immuni.2006.03.013
  152. 152. Koda T, Okuno T, Takata K, Honorat JA, Kinoshita M, Tada S, et al. Sema4A inhibits the therapeutic effect of IFN-β in EAE. Journal of Neuroimmunology. 2014;268:43-49. DOI: 10.1016/j.jneuroim.2013.12.014
  153. 153. Okuno T, Nakatsuji Y, Moriya M, Takamatsu H, Nojima S, Takegahara N, et al. Roles of Sema4D-plexin-B1 interactions in the central nervous system for pathogenesis of experimental autoimmune encephalomyelitis. Journal of Immunology. 2010;184:1499-1506. DOI: 10.4049/jimmunol.0903302
  154. 154. Comabella M, Fernández M, Martin R, Rivera-Vallvé S, Borrás E, Chiva C, et al. Cerebrospinal fluid chitinase 3-like 1 levels are associated with conversion to multiple sclerosis. Brain. 2010;133:1082-1093. DOI: 10.1093/brain/awq035
  155. 155. Cantó E, Tintoré M, Villar LM, Borrás E, Alvarez-Cermeño JC, Chiva C, et al. Validation of semaphorin 7A and ala-β-his-dipeptidase as biomarkers associated with the conversion from clinically isolated syndrome to multiple sclerosis. Journal of Neuroinflammation. 2014;11:181. DOI: 10.1186/s12974-014-0181-8
  156. 156. Gutiérrez-Franco A, Costa C, Eixarch H, Castillo M, Medina-Rodríguez EM, Bribián A, et al. Differential expression of sema3A and sema7A in a murine model of multiple sclerosis: Implications for a therapeutic design. Clinical Immunology. 2016;163:22-33. DOI: 10.1016/j.clim.2015.12.005
  157. 157. Fisher TL, Reilly CA, Winter LA, Pandina T, Jonason A, Scrivens M, et al. Generation and preclinical characterization of an antibody specific for SEMA4D. mAbs Journal. 2016;8:150-162. DOI: 10.1080/19420862.2015.1102813
  158. 158. LaGanke C, Samkoff L, Edwards K, Jung Henson L, Repovic P, Lynch S, et al. Safety/tolerability of the anti-semaphorin 4D antibody VX15/2503 in a randomized phase 1 trial. Neurology: Neuroimmunology & Neuroinflammation. 2017;4:e367. DOI: 10.1212/NXI.0000000000000367

Written By

Sebastian Quintremil, Fernando Medina Ferrer, Javier Puente, María Elsa Pando and María Antonieta Valenzuela

Submitted: 10 August 2018 Reviewed: 15 October 2018 Published: 10 December 2018