Next Article in Journal
Arthritogenic Alphavirus Capsid Protein
Next Article in Special Issue
Oligodendroglial Energy Metabolism and (re)Myelination
Previous Article in Journal
The Cellular Senescence Stress Response in Post-Mitotic Brain Cells: Cell Survival at the Expense of Tissue Degeneration
Previous Article in Special Issue
Neuron–Oligodendrocyte Communication in Myelination of Cortical GABAergic Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Interaction between Neurons and the Oligodendroglial Lineage in Multiple Sclerosis and Its Preclinical Models

by
Vasiliki Pantazou
1,2,
Thomas Roux
1,3,
Vanessa Oliveira Moreira
1,
Catherine Lubetzki
1,3,† and
Anne Desmazières
1,*,†
1
Paris Brain Institute (ICM), Sorbonne Université, CNRS, Inserm, GH Pitié-Salpêtrière, 47 boulevard de l’Hôpital, 75013 Paris, France
2
Service de Neurologie, Centre Hospitalier Universitaire Vaudois, 46 Rue du Bugnon, 1011 Lausanne, Switzerland
3
Assistance Publique-Hôpitaux de Paris, Neurology Department, Pitié Salpêtrière University Hospital, 75013 Paris, France
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 3 February 2021 / Revised: 6 March 2021 / Accepted: 7 March 2021 / Published: 11 March 2021
(This article belongs to the Special Issue Myelin and Oligodendrocyte-Neuron Interactions)

Abstract

:
Multiple sclerosis (MS) is a complex central nervous system inflammatory disease leading to demyelination and associated functional deficits. Though endogenous remyelination exists, it is only partial and, with time, patients can enter a progressive phase of the disease, with neurodegeneration as a hallmark. Though major therapeutic advances have been made, with immunotherapies reducing relapse rate during the inflammatory phase of MS, there is presently no therapy available which significantly impacts disease progression. Remyelination has been shown to favor neuroprotection, and it is thus of major importance to better understand remyelination mechanisms in order to promote them and hence preserve neurons. A crucial point is how this process is regulated through the neuronal crosstalk with the oligodendroglial lineage. In this review, we present the current knowledge on neuron interaction with the oligodendroglial lineage, in physiological context as well as in MS and its experimental models. We further discuss the therapeutic possibilities resulting from this research field, which might allow to support remyelination and neuroprotection and thus limit MS progression.

1. Introduction

Along with species evolution, there is an increasing complexity in the central nervous system (CNS) organization with, in particular, the acquisition of myelination [1,2], allowing for the efficient conduction of the nerve influx along myelinated fibers [3]. The choice of the axon to be myelinated, the thickness and length of the myelinated internode—even the pattern of myelin deposition along the axon—are not randomly selected but are a rather complex, carefully orchestrated collaboration of multiple cell types, with the two main actors in the CNS being the neuron and its myelinating cell—the oligodendrocyte (OL).
OLs enwrap axons with a multilayered lipid-rich membrane, the myelin. Myelin ensures electrical insulation allowing fast saltatory conduction along the axon [4], but also provides trophic and energetic support through delivery of lactate and pyruvate by oligodendrocytes [5,6]. In vitro and in vivo studies have suggested that a single mature OL can produce 5000 µm2 of myelin per day [7] and myelinate about 15 internodes in zebrafish [8] and up to 60 in rodents [9,10]. Establishing stable myelin sheaths can take from 2–3 h in the zebrafish [8] to a few days in rodents [11].
Although myelination can occur in vitro in the absence of axons—even around inert fibers such as electron-spun nanofibers [6,12]—neurons modulate myelination in vivo, where selective myelination has been observed among axons [13]. A typical example is the presence of the myelinated descending motor neurons in the medial forebrain in close vicinity to the not myelinated nigrostriatal dopaminergic neurons [14]. Axonal diameter is a key determinant to whether or not an axon will be myelinated with axons over 0.4 μm being preferentially myelinated in the CNS [15,16,17]. Numerous studies show a role of neuronal activity, growth factors and various signaling pathways in myelination regulation [18,19,20,21,22,23,24,25], although many aspects of the relationship between neurons and the oligodendroglial lineage are not fully elucidated yet.
This constant dialog between the CNS myelinating lineage and neurons is not only crucial for the development of the nervous system but is maintained throughout life and disease. A clear example of where this dialogue can be impaired is multiple sclerosis (MS), the most frequent demyelinating, inflammatory and neurodegenerative disease of the human central nervous system (CNS). MS affects more than two million people worldwide and is the leading cause of clinical disability and long-term neurological handicap amongst young adults [26,27]. In MS, myelin and/or the OL lineage are the target of an immune-mediated injury characterized by attacks disseminated throughout the CNS spatially and temporally [28]. As a result of these attacks, focal demyelination and neuronal damage are observed early in the disease course, but various degrees of remyelination do occur and allow for improved functional outcome [29,30,31,32,33,34].
In the present review, we summarize the current knowledge on neuron–OL interaction, in both demyelinating and remyelinating context in MS and its experimental models, concluding on the therapeutic possibilities resulting from this research field.

2. The Organization and Functioning of the Myelinated Axon

The myelinated axon is organized in internodes—axonal segments electrically insulated by myelin—alternating with small unmyelinated excitable domains enriched in voltage-gated sodium and potassium channels, called the nodes of Ranvier [35,36]. The ratio of the axonal diameter to the diameter of the myelin sheath added to the axon is called g-ratio and is used as an index of axonal myelination. Myelin sheath decreases membrane capacitance and increases membrane resistance. By doing so, it stabilizes the ionic charge, as the axolemma is depolarized during action potential regeneration and propagation at the nodes [4,21,37,38]. This leads to the so called saltatory conduction along myelinated axons, a function acquired along the evolution of species, allowing for higher velocities [1,39].
Recent studies suggest that besides this rather passive role on signal conduction, OLs actively participate in shaping electrical transmission [40,41]. Axonal proteins also regulate myelin deposition along the axon. Examples of such modulation include inhibition of myelin ensheathment by PSA-NCAM [42]. Furthermore, genetic deletion of some molecules implicated in paranodal and internodal axo-glial junctions, such as Cadm4 and Caspr, lead to the formation of axons with multiple layers of myelin sheaths, loose paranodal junctions or internodes of altered length [43,44].
Besides rapid conduction at the single neuron level, differences in myelin segment length and thickness may allow to create a synchronized network activity translating to higher cognitive functions [18,45,46,47,48,49,50]. Evolution of myelin segment characteristics and pattern have further been observed within hours following motor training or social isolation indicating that myelin—much like neurons—is a plastic structure able to adapt brain function to environmental stimuli [45,51,52,53].

3. Neuronal Activity as a Mediator of Myelin Plasticity

Even prior to becoming myelinating cells, OPCs sense their environment by monitoring electrical activity of surrounding neurons through NMDA, AMPA/kainate and GABA receptors [20,54,55,56,57,58]. Accumulating evidence suggests that OPCs maintain the capacity to communicate with neighboring neurons through non-synaptic junctions during different stages of development and maturation to myelinating oligodendrocytes [24,56,58] (for review see [59]). Examples of such a communication are the expression of NMDA receptors in mature OL, binding axon-released glutamate in the optic nerve of the rat [23], or the Ca2+ signals modulating myelin sheath length during myelination in zebrafish larvae [60].
Why is it important for the oligodendroglial lineage to maintain the capacity to monitor neuronal activity? Early in vitro experiments demonstrated that altering the sensing of neuronal activity interfered with myelination by using neurotoxins such as tetrodotoxin or α−scorpio toxin, which specifically either block or increase the probability of opening of voltage-gated sodium channels, respectively, Demerens et al. demonstrated that myelination was inhibited in electrically silent axons (tetrodotoxin) or enhanced by increasing neuronal activity (α−scorpio toxin) [19]. Later in vitro experiments in dorsal root ganglion neurons by Wake et al. showed that OPCs can detect axonal spiking and glutamate vesicular release and differentiate into myelinating OLs [56]. Further proof that neuronal activity stimulates OPC differentiation came more recently with the use of optogenetics in transgenic mice with myelination occurring preferentially on electrically active axons [21,61]. The development of two photon microscopy has allowed for in vivo live imaging, further expanding our knowledge. In zebrafish larvae, Mensch et al. have in particular shown that—similarly to in vitro experiments—individual OLs are generally capable to detect electrical activity through vesicular release by their neighboring neurons and, as a response, modulate their myelinating capacity [25].

4. Oligodendrocyte Lineage and Neuron: A Story of Mutual Dependence

Myelination has a dramatic impact on neuronal function and organization, as the mere act of myelin ensheathment around an axon improves the chances of survival of both cells. Indeed, OLs that do not manage to ensheath axons degenerate [18,62]. On the other hand, loss of OLs disrupts the normal nodal and paranodal architecture and leads to severe axonal pathology [50]. Furthermore, OLs promote neuronal survival though secretion of neurotrophic factors such as IGF-1 and GDNF [63] and provide metabolic support for the axon through delivery of lactate [64]. OLs are indeed highly enriched in monocarboxylase transporter 1 (MCT-1), the most abundant lactate transporter in CNS [17,65]. Lactate fuels axonal mitochondria, and by doing so, crucially supports highly energy-demanding axonal transport. Highlighting the necessity of this interaction for axonal integrity, MCT-1 knockout mice show age-dependent neuronal loss without evident demyelination [5]. Most recently, in an in vitro culture of mouse corpus callosum slices, Meyer et al. demonstrated that OLs fueled neuronal activity through glucose rather than lactate delivery [66], suggesting that OL–neuronal interactions are adapted to their local environment as they vary among cell types, locations and energetic needs. Interruption of this axo-glial metabolic support leads to deficits in information processing, even in the absence of demyelination, as demonstrated in the elegant experiments conducted by Moore et al. [67].

5. Altered Myelin and Axonal Damage: Lessons from the Animal Models

Studies in animal models have provided compelling evidence that axonal damage can result from an altered myelin. Myelin-associated glycoprotein (MAG) is one of the four major myelin proteins (along with myelin proteolipid protein—PLP, myelin basic protein—MBP and 2′,3′-cyclic-nucleotide phosphodiesterase—CNPase). MAG is expressed in the inner layer of myelin sheath and binds to axonal gangliosides GD1a and GT1b, establishing the axo-glial contact. In MAG-deficient mice, both motor development and myelin content appear normal. In closer inspection, subtle changes are observed, such as in neurofilament phosphorylation and spacing, and these changes ultimately lead to axonal degeneration in old age [68].
In PLP-deficient mice, myelin is still produced and packed thickly but is highly unstable. As a result, these mice present with normal early development but after 6–8 weeks they show focal axonal swellings containing membranous organelles, multivesicular bodies and mitochondrial dysfunction. By 18 months of age, they have significant motor impairment and extensive axonal degeneration [69]. Similar findings are reported in some patients with PLP mutations (Pelizaeus Merzbacher disease), who show CNS white matter disease, with axonal degeneration as measured by proton magnetic resonance spectroscopy analysis of N-acetyl-aspartate (NAA) [70].
On the other side of the spectrum, mutations of MBP result in unmyelinated axons or thin non-compacted myelin sheaths. Life span is reduced in MBP mutant rodents (up to 4 months for the Shiverer mice and up to 9 months for the Evans Shaker rats) [71,72]; nonetheless, even in terminal stages, demyelination is not followed by axonal damage. In these models, upregulation of neurotrophic factor expression in the remaining OLs and astrocytes is found [73], suggesting that degeneration may be prevented by compensatory mechanisms through axo-glial communication.

6. Myelin Insult and Neuronal Damage in Multiple Sclerosis

Focal demyelination of the CNS white matter is the hallmark of MS, as described by Charcot over 150 years ago [74,75]. Regardless of the perspective from which it is viewed (disease pathology, clinical and radiological findings, histopathological findings, outcomes), MS remains a highly complex disease. Clinically, the majority of patients will initially present with a single, self-limited, neurological episode, described as a clinically isolated syndrome, that typically involves the optic nerve, brainstem, or spinal cord. MS diagnosis is possible even at this early stage in the presence of typical findings on an abnormal MRI scan and/or positive oligoclonal bands in the cerebrospinal fluid [76].
In MS, 85% of the patients present with a relapsing remitting form of the disease [27]. Active demyelination is considered the histopathological equivalent of the clinical relapse [77]. Four patterns of active lesions have been described (pattern I–IV) [78], all sharing demyelination as a common denominator with various degrees of inflammatory cell infiltration (peripheral macrophages, activated microglia, T lymphocytes, complement, plasma cells) and OL loss. OL loss or myelin destruction will lead to a conduction block, which can translate to a neurological deficit, and if prolongated, will also lead to neuronal damage [79] (Figure 1). Depending on the number of adjacent internodes being demyelinated, nerve conduction velocity can be decreased or completely blocked. So far and to the best of our knowledge, how many internodes need to be demyelinated to cause a conduction block still remains to be established.
However, axonal loss and demyelination in absence of macrophage infiltration can be found in about 15% of patients who develop progressive onset MS described as primary progressive MS (PPMS) [27,77]. Furthermore, focal changes have been documented in normal-appearing white matter of MS months to years before the appearance of MRI gadolinium-enhanced lesions. Altered axonal organization with changes in the distribution of sodium and potassium channels, early alterations of paranodal structures [35,80,81,82,83], focal axonal swelling or degeneration can be found prior to OL loss [84]. Such findings opened the discussion on whether axonal degeneration is a direct consequence of demyelination, an independent mechanism or even the prime event leading to demyelination.

7. Endogenous Remyelination Occurs in MS and Its Animal Models

Alongside with demyelinated axons, lesions with evidence of remyelination, also known as “shadow plaques”, are characterized by the presence of thinner myelin sheaths and shorter internodes. Post-mortem human MS tissue studies have shown that 20% of the patients showed fully remyelinated lesions [33]. For a brief history on discovery of remyelination and discussion on whether shadow plaques are signs of remyelination or, on the contrary, of ongoing demyelination, see [31,85]. Shorter internodal segments, absence of immune cell infiltration and only few microglia/macrophages within the lesion help to distinguish shadow plaques from early demyelinating lesions.
Although often extensive in animal models of demyelination, the degree to which remyelination occurs in MS patients is variable. It differs among locations, with a higher degree of remyelination in cortical and subcortical lesions compared to periventricular or cerebellar lesion [86], and among patients, as shown in histopathological studies and PET-MRI studies using myelin markers ([11C]PIB) [32,33,87,88]. Subgroups of MS patients with either extensive or limited remyelination with variable myelin content are identified, suggesting that inter-individual myelination capacity may further influence the ability of lesions to remyelinate.
Whether remyelination occurs from surviving OLs or newly differentiated OLs derived from OPCs have been studied over the last two decades and is comprehensively discussed in a recent review by [89]. The notion that remyelination requires OPC recruitment and differentiation comes from multiple observations: 1. After OL and OPC depletion by ethidium bromide induced demyelination in rodents, OPCs appear in the lesion site prior to mature OLs [90]; 2. Transitioning phenotypes of OPCs have been observed in demyelinating lesions in adult rat brain after toxic-induced demyelination. Different surface or nuclear markers can help differentiate OL lineage cells, as the proteoglycan NG2, the platelet derived growth factor receptor α (PDGFRα), OLIG1, OLIG2 and Nkx2.2 are expressed early in OPC development, while O4 and glycolipid galactocerebrosidase (GalC) appear later in immature or pre-myelinating OL and so on [91,92]; 3. After single cell OL laser ablation, an homeostatic replacement takes place by a newly differentiated OPC who produces myelin with thinner and shorter internodal sheaths, as observed by live two photon imaging, but sufficient to restore function [93]; 4. Newly differentiated OLs in a capsaicin-induced model of demyelination in zebrafish were more efficient in myelin sheath formation than OL survivors [94]; 5. In post-mortem analysis of MS brains, there was an increased density of OPCs in the subventricular zones compared to controls, which would further increase in active lesions compared to normal appearing white matter (NAWM) [95]. Furthermore, in the cuprizone model, experimental conditions favoring differentiation of subventricular zone neural stem cells into oligodendrocytes resulted in remyelination with normal thickness myelin sheath restoring normal nerve conduction [96,97].
The role of mature OLs in remyelination has been debated, since these cells do not seem to have migrating nor proliferating capacities. Nonetheless, almost 30 years ago, this notion was challenged when mature OLs grafted into newborn shiverer mouse brain produced new myelin sheaths around denuded axons [98]. The use of multiphoton imaging to track OL lineage response following cuprizone-induced demyelination in mice cerebral cortex [11] and capsaicin-induced demyelination in zebrafish [94] have provided more insight into myelin regeneration. In both studies, newly differentiated OLs produced large number of myelin sheaths forming internodes along the axon. Surviving OLs could only occasionally produce new myelin sheaths, often misplacing them to neuronal soma [94]. Although this process was less efficient, it occurred over protracted periods of time and was enhanced by training [11]. These observations in the animal models suggest that even though both mature and newly differentiated OLs can produce myelin sheaths, it is newly differentiated OLs, rather than mature OLs, that orchestrate remyelination in adult brain following injury in these models.
In humans, given the low level of oligodendrocyte generation in adult CNS tissue [99,100], it is unlikely that newly generated OLs alone can respond to the demand in myelin production. Recent works have indeed shown a very low rate of new OL generation in shadow plaques of MS patients [100], further favoring a role of pre-existing OLs for remyelination in MS. However, remyelination by mature OLs happens only at low levels in mice [11,94], underlying a potential difference in remyelination mechanism in MS and its animal models.

8. Remyelination Failure in MS

Even though remyelination is possible in MS, the majority of lesions show only partial remyelination or no remyelination at all [33], despite sufficient number of OL-lineage cells were present at the lesion sites [101], suggesting that other mechanisms than paucity of myelinating cells participate in successful remyelination. Repetitive rounds of demyelination-remyelination is a characteristic feature of the disease in humans [102]. Using lesion segmentation from MTR images, MRI studies have demonstrated recurrent rounds of inflammatory demyelination in pre-existing lesions, associated with lower capacities of recovery [79,88]. Furthermore, ageing has been associated with reduced ability to remyelinate [103,104,105]. Although mechanisms that lead to remyelination failure are not yet fully elucidated, what has been made obvious is that remyelination failure correlates with increased clinical disability and extensive axonal loss (Figure 1) [30,103]. This paradigm is seen in 15–30% of RRMS patients, who, over time, develop progressive disability described as secondary-progressive MS [27]. These patients show extensive axonal degeneration and little evidence of remyelination, although a subset of them may still exhibit inflammatory activity as shown clinically by relapses, and by areas with sharp limits of demyelination and a rim of macrophages at the lesion edge in pathology [77], that mediate active ongoing myelin breakdown and axonal damage [88]. The rim is also identifiable in MRI studies as a persistent phase rim [106] or an iron rim at the lesion edge [107], which can provide a useful prognostic tool in MS patients [108].
Furthermore, in remyelinated lesions, axonal density is higher with a less extensive degree of axonal damage than in non-remyelinated lesions (as measured by β-amyloid-precursor protein staining), suggesting a positive effect of new myelin sheath formation to axonal preservation [109]. Similar findings are reported in experimental autoimmune encephalomyelitis (EAE), an inflammatory animal model of MS, in which remyelination confers preservation of axonal density and improvement of functional recovery and outcome [110].
It is thus essential to better understand how to promote an adequate neuro-glial crosstalk to support remyelination and limit the progression of the disease.

9. After a Demyelinating Insult, What Are the Required Steps for New Myelin Sheath Formation?

In order to promote repair in MS, it would be of interest to increase OPCs recruitment to the lesions and favor their differentiation in myelinating OLs, as well as promote remyelination by surviving OLs. Indeed, the prevailing opinion is that efficient myelin repair requires activation and recruitment of OPCs to demyelinated areas, followed by their differentiation into mature OLs and target recognition of the axon to remyelinate (Figure 2). At the lesion site, newly differentiated OLs along with surviving OLs proceed to new myelin sheath formation and production of stable myelin sheaths [30,89]. As discussed below, each of these steps requires the interaction between neurons and the oligodendroglial lineage.
Following activation, OPCs migrate to the lesion site attracted by guidance cues [47,111]. Growth factors like insulin-like growth factor 1 (IGF-1) and glial cell line-derived neurotrophic factor (GDNF) have also been shown to act in OPC recruitment [47,89]. Among attractant cues, Semaphorin 3F has been identified both in in vitro and in vivo models. In contrast, Netrin was shown to act as a repellent for OPCs [112], as well as Semaphorin 3A [111]. Neurons further secrete adenosine and stimulate neighbor astrocytes to produce leukemia inhibitory factor or CNTF to promote OPC differentiation into myelinating OLs [113,114]. The next step to remyelination is careful selection of the axon to remyelinate [18,62]. As previously stated, neuronal activity also plays a role in remyelination [20,61]. OPCs are not only sensitive to the presence of neuronal activity, but also to the pattern of activity, which modulates differently their proliferation and differentiation [55]. Furthermore, following a demyelinating insult, dispersion of voltage-gated sodium channels is observed along the axon [80,81,115]. In a very recent study, it has been shown, using a cuprizone-induced demyelination model, that βIV spectrin, which forms a complex with Ankyrin-G to link voltage-gated channels to actin cytoskeleton at the nodes of Ranvier, remains however clustered after demyelination in the cortical layers of the somatosensory cortex of adult mice, which could provide reclustering cues for nodal voltage-gated ionic channels [116]. Interestingly, multiple studies conducted by us and others reported nodal reclustering can occur prior to (re)myelination and suggest they could impact conduction velocity and myelin initiation guidance [35,117,118,119].
In contrast, pathways inhibiting OPC maturation and recruitment have also been identified such as axonal expressed Jagged-1. Jagged-1 protein binds OPC expressed NOTCH-1, inhibiting their differentiation [119]. More recently, the Wnt pathway and the mechano-responsive ion channel Piezo1 activation have also been identified as OPC differentiation inhibitors, possibly involved in age-related remyelination capacity decline [30,89].
Moreover, the perineuronal extracellular matrix (ECM) could also modulate remyelination abilities in MS. Indeed, hyaluronan, as well as chondroitin sulfate proteoglycans (CSPGs) are described to accumulate at the border of actively demyelinating plaques, where they inhibit the migration and differentiation of oligodendrocyte precursor cells (OPCs) and remyelination [120]. Enhanced OPC differentiation and remyelination were further observed in cuprizone and EAE mice models when suppressing CSPGs activity [121,122].
Better understanding the complex dialogue between neurons and the oligodendroglial lineage and how it is balanced in the context of their direct environment will thus be required to adequately promote axon remyelination in MS.

10. Other Cellular Contributors to Successful Remyelination

It is now clear from the abundant literature on the subject that the process of myelin repair requires a carefully orchestrated collaboration between multiple players (Figure 2).
Whether astrocytes bring more damage or inhibit repair is still under debate [123], although evidence from studies using ethidium bromide-induced demyelination speaks favorably for their role during remyelination, as elimination of astrocytes resulted in absence of CNS remyelination [124]. In addition, phagocytosis of myelin debris by surrounding microglia/macrophages promotes remyelination [125,126,127]. Their role further extends beyond phagocytosis, as microglial switch to a pro-regenerative profile promotes remyelination, and maintenance of a more pro-inflammatory phenotype is deleterious for repair. Microglia secrete various factors that can modulate the oligodendroglial lineage and promote (re)myelination [128,129,130,131] (Figure 2). The CXCL12 chemokine is expressed by microglia in MS lesions and has been described to modulate OPCs chemoattraction and differentiation [132]. Microglia is also described as a major source of iron for OPCs during developmental myelination, promoting their proliferation and differentiation and could further play a role in remyelination [133]. Furthermore, both astrocytes and microglia can perceive neuronal activity, suggesting they could participate in an indirect dialogue between neurons and the oligodendroglial lineage (for review, see [59]).
It has further been shown that failure in OPCs recruitment and differentiation increases with age [105,134,135], and the Franklin laboratory showed this could be reverted using a heterochronic parabiosis experiment, in which old mice shared systemic circulation with younger mice. After LPC induction of spinal cord demyelination, they observed a higher rate of remyelination in these mice comparing to the isochronic controls (old mice sharing systemic circulation with their age homologues), due to the recruitment of “younger” blood-derived monocytes from their partners. Moreover, the importance of monocyte/macrophage recruitment in remyelination was highlighted using mice deficient in CCR2, a major player in macrophage recruitment, in which a significant decrease of proliferating OPC was observed when compared to controls [104].

11. Myelin Repair: From Animal Models to Human Translation

11.1. The Challenges of Myelin Repair Strategies.

The past two decades have been an era of exponential growth of disease-modifying immunotherapies acting on the inflammatory component of the disease, preventing these recurrent episodes of inflammation [27]. As a result, patients experience fewer or no relapses, show less or no gadolinium-enhanced lesions and less disease progression (described as NEDA-3 status) [136]. The recent epidemiological studies have shown that the rate of conversion to secondary progressive MS has been reduced from 50% to less than 30% over the last 10 years [27].
However, these therapies are still insufficient to block this progressive phase, and half of the patients will suffer a disability accumulation [27,30,137,138]. In the light of this unmet need, the scientific community is now focusing its efforts on ways to promote remyelination in an effort to ensure neuroprotection [139].
A key challenge will be that at any given time point of the disease, the patterns of demyelination and axonal degeneration are different among patients or even in different brain regions of a given patient [140]. This heterogeneity in MS brain with variable—but without a doubt concomitant—presence of OL, OPC and axonal dysfunction from early on, defines the therapeutic challenge of the disease. Regarding this aspect, a better understanding of the basic mechanisms underlying MS pathology and repair is required.
In parallel, in the quest of remyelination candidate therapies, an important step is coming up with the development of screening strategies that are both adapted and available. In vitro assays of rodent-derived OPCs or micropillar arrays [141] provide a fast way to test the promyelinating effect of bioactive molecules but they offer little insight in the molecules’ physiological mechanism of action. Animal models including toxin or chemically-induced demyelination followed by spontaneous remyelination bypass the aforementioned problem. Among them, zebrafish larvae and Xenopus laevis, allow for a robust large screening platform, while EAE models have the advantage of resembling the closest to the human MS disease but are not suitable for the screening of large numbers of compounds (for review, see [30]).
Using these different approaches, a multitude of compounds has been tested following two strategies: revisiting existing drugs for their potential role on myelination or developing new molecule-targeting pathways of OPC differentiation and axo-glial interaction.

11.2. Repurposing Existing Drugs for Their Remyelinating Potential

A great example of a molecule repurposed for its promyelinating potential is clemastine fumarate. Clemastine is a first-generation histamine H1 receptor antagonist used since the early 1990s in the treatment of allergies. Clemastine crosses the blood–brain barrier and was found to promote OPC differentiation and myelination in a high-throughput in vitro screening and its efficiency in remyelination was confirmed in mouse in vivo following LPC-induced focal demyelination [141]. The molecule was then tested in the cuprizone-induced model of demyelination, with similar results [142]. Its effect on OPC maturation is thought to be mediated by an antimuscarinic action in the M1 muscarinic acetylcholine receptor Chrm1 [110]. From clinical biology to human pathology, clemastine was tested in chronic optic neuritis patients in a phase II trial [143] (ReBUILD trial) where it met its primary outcome and showed a small reduction in evoked potential latency but failed to improve clinical outcome. The drug is now tested in acute neuritis patients in the ReCOVER trial (NCT02521311).
Other than clemastine, a large number of molecules have been tested in preclinical studies for their potential benefit in myelination. Less than 20 have demonstrated significantly enhanced remyelination in animal models, either by promoting OPC differentiation or by reducing microglial and astrocyte activation [139]. Such drugs involve anticholinergics and antimuscarinics (GSK239512 [144], benzatropine [145]), antiemetics (domperidone [146], antipsychotic drugs (quetiapine [147]), selective estrogen modulators (progesterone, tamoxifen [148]), hormones (ACTH [149]), antifungal drugs (miconazole [150]), antidiabetic drugs (metformin [151]), immunomodulatory drugs (glatiramer acetate, siponimod, teriflunomide [151,152,153] and statins [154]. Simvastatin, a widely used statin, is one of the few of these molecules being tested in clinical trials (Table 1). In a phase II study including 70 SPMS patients (MS-STAT study), the authors report a 43% reduction of annual brain atrophy rate and a 63% reduction in the number of black holes with simvastatin treatment [155]. Less promising results were obtained in the phase IV SIMCOMBIN trial in RRMS patients, as an add-on therapy to interferon-beta, where simvastatin failed to further reduce annualized relapse rate [156] underlying a pro-regenerative rather than anti-inflammatory role of the drug.
On the other hand, not all molecules tested had a null or positive impact on myelination. Some seemed to hasten remyelination after a demyelinating insult, such as valproic acid, a first generation antiepileptic drug [157].

11.3. Strategies Promoting OPC Differentiation

The second approach involves molecules that have been designed to target a specific pathway in OPC differentiation or maturation. Multiple monoclonal antibodies have been developed the last two decades and are being tested in clinical studies. Studies showed that Semaphorin 4D is upregulated in OPCs upon spinal cord injury and inhibits remyelination in EAE [158,159], while using an anti-Semaphorin 4D rescues oligodendroglial differentiation, with a strong reduction of inflammatory activity in EAE. A humanized monoclonal anti-Semaphorin 4D antibody (VX15/2503) was tested in a phase I study with no major side effects [160].
Similar “from bench to clinical translation” examples include temelimab/GNbAC1, a humanized antibody directed against the envelope protein (ENV) of the human endogenous retrovirus (HERV) [161] and opicinumab, an anti-LINGO-1 monoclonal antibody [162]. Following observations in in vitro OPC cultures that ENV protein inhibited OPC differentiation via activation of Toll-like receptor 4 [163], temelimab testing completed successfully a phase I trial in 21 individuals. As a consequence, the phase IIa CHANGE-MS trial [161] and phase IIb ANGEL-MS trial (Hartung et al., presented in ECTRIMS 2019) investigated efficacy of variable regimens of temelimab on over 200 MS patients. The authors report a reduction in cortical and thalamic atrophy and a benefit in magnetization transfer ratio (MTR) in both NAWM and cerebral cortical bands, suggesting an effect on remyelination. However, although loss of LINGO-1 enhanced myelin sheath formation and myelination in preclinical studies [162] and showed a small efficacy in reducing evoked potential latency in patients with acute optic neuritis (RENEW trial) [164], two follow up phase II studies (SYNERGY, AFFINITY) failed to reach their primary endpoints (Biogen Press Communication in October 2020) [165].
Inhibition of Bruton’s tyrosine kinase (BTK), a member of the Tec family of kinases, expressed by cells of hemopoietic origin and microglia has moreover been shown to favor remyelination [166] and a phase II clinical trial in MS using a BTK inhibitor has shown beneficial outcome [167]. As a result, four phase III trials (GEMINI I & II, PERSEUS, HERCULES) are ongoing examining BTKs’ efficacy in different MS populations (relapsing, primary and secondary progressive).
More recently, new approaches directed against the immune component in relation to its impact on OL lineage have emerged. Among them bexarotene/IRX4204, a specific agonist of retinoic acid receptor gamma (RXR-γ), reduced EAE severity by decreasing Th17 responses and boosting OPC differentiation [168]. The results of the phase IIa study including 52 active MS patients randomized to receive either bexarotene or placebo were recently communicated (ECTRIMS, September 2020). The investigators reported an advantage of bexarotene over placebo in reducing mean magnetization transfer ratio in a subset of MRI lesions, although drug-treated patients experienced serious side effects.
Furthermore, the use of anacardic acid improved the clinical scores of both cuprizone-induced demyelination and EAE through induction of IL-33 and upregulation of genes involved in myelin protein synthesis [169]. Melero-Jerez et al. very recently showed in the EAE model that myeloid-derived suppressor cells directly influence OPC survival, proliferation, and differentiation through upregulation of osteopontin [170].
These data highlight a crucial pathogenic interaction between innate immunity and the CNS myelination potential, opening new ways to promote effective myelin preservation and repair in MS patients. Along with the results of the ongoing studies, future directions include translation to clinical trials for the molecules showing remyelinating potency in the animal models of demyelination, as well as potential for cell grafts promoting remyelination, including fascinating approaches such as administration of encapsuled OPCs within hydrogel particles via the nose-to-brain pathway [171].

12. Conclusions

Even if a solid conclusion from our current knowledge on remyelination strategies is premature, the results of this research field are exciting for therapeutic development in multiple sclerosis. Understanding the complex interaction of neurons with the oligodendroglial lineage throughout development and disease and how it is modulated by surrounding glial cells may be the key determinant to successfully promote remyelination, ensure neuroprotection and reduce disability progression, which is the main unmet need in multiple sclerosis.

Author Contributions

Conceptualization, V.P., T.R. and A.D.; writing—original draft preparation, V.P., T.R, V.O.M. and A.D.; writing—review and editing, V.P., T.R., C.L. and A.D.; funding acquisition, A.D., V.P., T.R., C.L. and A.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by ARSEP R20218DD grant (to A.D.), Brain Paris Institute and INSERM; V.P. is supported by the “Funds for research”, University of Lausanne, Switzerland; V.O.M. is supported by ANR JC R17127DD grant (to A.D.). C.L. received support from Biogen (Ph.D. student salary).

Acknowledgments

We thank Nathalie Sol-Foulon for her invitation to participate in this topic and Bernard Zalc for his careful reading of the manuscript.

Conflicts of Interest

The other authors declare no conflict of interest. Figures were created with Biorender.com accessed on 7 March 2021.

References

  1. Zalc, B.; Goujet, D.; Colman, D. The Origin of the Myelination Program in Vertebrates. Curr. Biol. 2008, 18, R511–R512. [Google Scholar] [CrossRef] [Green Version]
  2. Zalc, B.; Colman, D.R. Origins of Vertebrate Success. Science 2000, 288, 271–272. [Google Scholar] [CrossRef]
  3. Rushton, W.A.H. A Theory of the Effects of Fibre Size in Medullated Nerve. J. Physiol. 1951, 115, 101–122. [Google Scholar] [CrossRef] [PubMed]
  4. Waxman, S.G.; Bennett, M.V.L. Relative Conduction Velocities of Small Myelinated and Non-Myelinated Fibres in the Central Nervous System. Nat. New Biol. 1972, 238, 217–219. [Google Scholar] [CrossRef]
  5. Fünfschilling, U.; Supplie, L.M.; Mahad, D.; Boretius, S.; Saab, A.S.; Edgar, J.; Brinkmann, B.G.; Kassmann, C.M.; Tzvetanova, I.D.; Möbius, W.; et al. Glycolytic Oligodendrocytes Maintain Myelin and Long-Term Axonal Integrity. Nature 2012, 485, 517–521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Lee, S.; Leach, M.K.; Redmond, S.A.; Chong, S.Y.C.; Mellon, S.H.; Tuck, S.J.; Feng, Z.-Q.; Corey, J.M.; Chan, J.R. A Culture System to Study Oligodendrocyte Myelination Processes Using Engineered Nanofibers. Nat. Methods 2012, 9, 917–922. [Google Scholar] [CrossRef] [Green Version]
  7. Pfeiffer, S.; Warrington, A.; Bansal, R. The Oligodendrocyte and Its Many Cellular Processes. Trends Cell Biol. 1993, 3, 191–197. [Google Scholar] [CrossRef]
  8. Czopka, T.; ffrench-Constant, C.; Lyons, D.A. Individual Oligodendrocytes Have Only a Few Hours in Which to Generate New Myelin Sheaths In Vivo. Dev. Cell 2013, 25, 599–609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Dumas, L.; Heitz-Marchaland, C.; Fouquet, S.; Suter, U.; Livet, J.; Moreau-Fauvarque, C.; Chédotal, A. Multicolor Analysis of Oligodendrocyte Morphology, Interactions, and Development with Brainbow: Multicolor Imaging of Myelination. Glia 2015, 63, 699–717. [Google Scholar] [CrossRef]
  10. Matthews, M.A.; Duncan, D. A Quantitative Study of Morphological Changes Accompanying the Initiation and Progress of Myelin Production in the Dorsal Funiculus of the Rat Spinal Cord. J. Comp. Neurol. 1971, 142, 1–22. [Google Scholar] [CrossRef]
  11. Bacmeister, C.M.; Barr, H.J.; McClain, C.R.; Thornton, M.A.; Nettles, D.; Welle, C.G.; Hughes, E.G. Motor Learning Promotes Remyelination via New and Surviving Oligodendrocytes. Nat. Neurosci. 2020, 23, 819–831. [Google Scholar] [CrossRef]
  12. Bechler, M.E.; Byrne, L.; ffrench-Constant, C. CNS Myelin Sheath Lengths Are an Intrinsic Property of Oligodendrocytes. Curr. Biol. 2015, 25, 2411–2416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Lubetzki, C.; Demerens, C.; Anglade, P.; Villarroya, H.; Frankfurter, A.; Lee, V.M.; Zalc, B. Even in Culture, Oligodendrocytes Myelinate Solely Axons. Proc. Natl. Acad. Sci. USA 1993, 90, 6820–6824. [Google Scholar] [CrossRef] [Green Version]
  14. Zalc, B. The Acquisition of Myelin: An Evolutionary Perspective. Brain Res. 2016, 1641, 4–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Goebbels, S.; Wieser, G.L.; Pieper, A.; Spitzer, S.; Weege, B.; Yan, K.; Edgar, J.M.; Yagensky, O.; Wichert, S.P.; Agarwal, A.; et al. A Neuronal PI(3,4,5)P3-Dependent Program of Oligodendrocyte Precursor Recruitment and Myelination. Nat. Neurosci. 2017, 20, 10–15. [Google Scholar] [CrossRef] [Green Version]
  16. Hildebrand, C.; Remahl, S.; Persson, H.; Bjartmar, C. Myelinated Nerve Fibres in the CNS. Prog. Neurobiol. 1993, 40, 319–384. [Google Scholar] [CrossRef]
  17. Lee, Y.; Morrison, B.M.; Li, Y.; Lengacher, S.; Farah, M.H.; Hoffman, P.N.; Liu, Y.; Tsingalia, A.; Jin, L.; Zhang, P.-W.; et al. Oligodendroglia Metabolically Support Axons and Contribute to Neurodegeneration. Nature 2012, 487, 443–448. [Google Scholar] [CrossRef] [PubMed]
  18. Barres, B.A.; Raff, M.C. Axonal Control of Oligodendrocyte Development. J. Cell Biol. 1999, 147, 1123–1128. [Google Scholar] [CrossRef]
  19. Demerens, C.; Stankoff, B.; Logak, M.; Anglade, P.; Allinquant, B.; Couraud, F.; Zalc, B.; Lubetzki, C. Induction of Myelination in the Central Nervous System by Electrical Activity. Proc. Natl. Acad. Sci. USA 1996, 93, 9887–9892. [Google Scholar] [CrossRef] [Green Version]
  20. Gautier, H.O.B.; Evans, K.A.; Volbracht, K.; James, R.; Sitnikov, S.; Lundgaard, I.; James, F.; Lao-Peregrin, C.; Reynolds, R.; Franklin, R.J.M.; et al. Neuronal Activity Regulates Remyelination via Glutamate Signalling to Oligodendrocyte Progenitors. Nat. Commun. 2015, 6, 8518. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Gibson, E.M.; Purger, D.; Mount, C.W.; Goldstein, A.K.; Lin, G.L.; Wood, L.S.; Inema, I.; Miller, S.E.; Bieri, G.; Zuchero, J.B.; et al. Neuronal Activity Promotes Oligodendrogenesis and Adaptive Myelination in the Mammalian Brain. Science 2014, 344, 1252304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Hines, J.H.; Ravanelli, A.M.; Schwindt, R.; Scott, E.K.; Appel, B. Neuronal Activity Biases Axon Selection for Myelination in Vivo. Nat. Neurosci. 2015, 18, 683–689. [Google Scholar] [CrossRef] [Green Version]
  23. Káradóttir, R.; Cavelier, P.; Bergersen, L.H.; Attwell, D. NMDA Receptors Are Expressed in Oligodendrocytes and Activated in Ischaemia. Nature 2005, 438, 1162–1166. [Google Scholar] [CrossRef] [Green Version]
  24. Kukley, M.; Nishiyama, A.; Dietrich, D. The Fate of Synaptic Input to NG2 Glial Cells: Neurons Specifically Downregulate Transmitter Release onto Differentiating Oligodendroglial Cells. J. Neurosci. 2010, 30, 8320–8331. [Google Scholar] [CrossRef]
  25. Mensch, S.; Baraban, M.; Almeida, R.; Czopka, T.; Ausborn, J.; El Manira, A.; Lyons, D.A. Synaptic Vesicle Release Regulates Myelin Sheath Number of Individual Oligodendrocytes in Vivo. Nat. Neurosci. 2015, 18, 628–630. [Google Scholar] [CrossRef] [PubMed]
  26. Noseworthy, J.H.; Lucchinetti, C.; Rodriguez, M.; Weinshenker, B.G. Multiple Sclerosis. N. Engl. J. Med. 2000, 343, 938–952. [Google Scholar] [CrossRef] [PubMed]
  27. Thompson, A.J.; Baranzini, S.E.; Geurts, J.; Hemmer, B.; Ciccarelli, O. Multiple Sclerosis. Lancet 2018, 391, 1622–1636. [Google Scholar] [CrossRef]
  28. Lucchinetti, C.F.; Brück, W.; Rodriguez, M.; Lassmann, H. Distinct Patterns of Multiple Sclerosis Pathology Indicates Heterogeneity in Pathogenesis. Brain Pathol. 1996, 6, 259–274. [Google Scholar] [CrossRef] [PubMed]
  29. Franklin, R.J.M. Regenerating CNS Myelin—from Mechanisms to Experimental Medicines. Nat. Rev. Neurosci. 2017, 18, 753–769. [Google Scholar] [CrossRef]
  30. Lubetzki, C.; Zalc, B.; Williams, A.; Stadelmann, C.; Stankoff, B. Remyelination in Multiple Sclerosis: From Basic Science to Clinical Translation. Lancet Neurol. 2020, 19, 678–688. [Google Scholar] [CrossRef]
  31. Neumann, B.; Foerster, S.; Zhao, C.; Bodini, B.; Reich, D.S.; Bergles, D.E.; Káradóttir, R.T.; Lubetzki, C.; Lairson, L.L.; Zalc, B.; et al. Problems and Pitfalls of Identifying Remyelination in Multiple Sclerosis. Cell Stem Cell 2020, 26, 617–619. [Google Scholar] [CrossRef]
  32. Bodini, B.; Veronese, M.; García-Lorenzo, D.; Battaglini, M.; Poirion, E.; Chardain, A.; Freeman, L.; Louapre, C.; Tchikviladze, M.; Papeix, C.; et al. Dynamic I Maging of I Ndividual R Emyelination P Rofiles in M Ultiple S Clerosis. Ann. Neurol. 2016, 79, 726–738. [Google Scholar] [CrossRef] [Green Version]
  33. Patrikios, P.; Stadelmann, C.; Kutzelnigg, A.; Rauschka, H.; Schmidbauer, M.; Laursen, H.; Sorensen, P.S.; Bruck, W.; Lucchinetti, C.; Lassmann, H. Remyelination Is Extensive in a Subset of Multiple Sclerosis Patients. Brain 2006, 129, 3165–3172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Albert, M.; Antel, J.; Brück, W.; Stadelmann, C. Extensive Cortical Remyelination in Patients with Chronic Multiple Sclerosis. Brain Pathol. 2007, 17, 129–138. [Google Scholar] [CrossRef]
  35. Lubetzki, C.; Sol-Foulon, N.; Desmazières, A. Nodes of Ranvier during Development and Repair in the CNS. Nat. Rev. Neurol. 2020, 16, 426–439. [Google Scholar] [CrossRef]
  36. Poliak, S.; Peles, E. The Local Differentiation of Myelinated Axons at Nodes of Ranvier. Nat. Rev. Neurosci. 2003, 4, 968–980. [Google Scholar] [CrossRef] [PubMed]
  37. Arancibia-Carcamo, I.L.; Attwell, D. The Node of Ranvier in CNS Pathology. Acta Neuropathol. 2014, 128, 161–175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Bakiri, Y.; Káradóttir, R.; Cossell, L.; Attwell, D. Morphological and Electrical Properties of Oligodendrocytes in the White Matter of the Corpus Callosum and Cerebellum: Oligodendrocyte Electrical Properties. J. Physiol. 2011, 589, 559–573. [Google Scholar] [CrossRef] [PubMed]
  39. Hartline, D.K.; Colman, D.R. Rapid Conduction and the Evolution of Giant Axons and Myelinated Fibers. Curr. Biol. 2007, 17, R29–R35. [Google Scholar] [CrossRef] [Green Version]
  40. Dubessy, A.-L.; Mazuir, E.; Rappeneau, Q.; Ou, S.; Abi Ghanem, C.; Piquand, K.; Aigrot, M.-S.; Thétiot, M.; Desmazières, A.; Chan, E.; et al. Role of a Contactin Multi-Molecular Complex Secreted by Oligodendrocytes in Nodal Protein Clustering in the CNS. Glia 2019, 67, 2248–2263. [Google Scholar] [CrossRef] [Green Version]
  41. Thetiot, M.; Freeman, S.A.; Roux, T.; Dubessy, A.; Aigrot, M.; Rappeneau, Q.; Lejeune, F.; Tailleur, J.; Sol-Foulon, N.; Lubetzki, C.; et al. An Alternative Mechanism of Early Nodal Clustering and Myelination Onset in GABAergic Neurons of the Central Nervous System. Glia 2020, 68, 1891–1909. [Google Scholar] [CrossRef] [PubMed]
  42. Charles, P. Re-Expression of PSA-NCAM by Demyelinated Axons: An Inhibitor of Remyelination in Multiple Sclerosis? Brain 2002, 125, 1972–1979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Elazar, N.; Vainshtein, A.; Rechav, K.; Tsoory, M.; Eshed-Eisenbach, Y.; Peles, E. Coordinated Internodal and Paranodal Adhesion Controls Accurate Myelination by Oligodendrocytes. J. Cell Biol. 2019, 218, 2887–2895. [Google Scholar] [CrossRef] [Green Version]
  44. Klingseisen, A.; Ristoiu, A.-M.; Kegel, L.; Sherman, D.L.; Rubio-Brotons, M.; Almeida, R.G.; Koudelka, S.; Benito-Kwiecinski, S.K.; Poole, R.J.; Brophy, P.J.; et al. Oligodendrocyte Neurofascin Independently Regulates Both Myelin Targeting and Sheath Growth in the CNS. Dev. Cell 2019, 51, 730–744.e6. [Google Scholar] [CrossRef] [Green Version]
  45. Fields, R.D. White Matter in Learning, Cognition and Psychiatric Disorders. Trends Neurosci. 2008, 31, 361–370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Itoh, T. Neuron-Oligodendrocyte Interaction in Neuroinflammation. Clin. Exp. Neuroimmunol. 2015, 6, 232–244. [Google Scholar] [CrossRef]
  47. Simons, M.; Trajkovic, K. Neuron-Glia Communication in the Control of Oligodendrocyte Function and Myelin Biogenesis. J. Cell Sci. 2006, 119, 4381–4389. [Google Scholar] [CrossRef] [Green Version]
  48. Karimian, M.; Dibenedetto, D.; Moerel, M.; Burwick, T.; Westra, R.L.; De Weerd, P.; Senden, M. Effects of Synaptic and Myelin Plasticity on Learning in a Network of Kuramoto Phase Oscillators. Chaos 2019, 29, 083122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Bells, S.; Lefebvre, J.; Longoni, G.; Narayanan, S.; Arnold, D.L.; Yeh, E.A.; Mabbott, D.J. White Matter Plasticity and Maturation in Human Cognition. Glia 2019, 67, 2020–2037. [Google Scholar] [CrossRef] [PubMed]
  50. Oluich, L.-J.; Stratton, J.A.S.; Lulu Xing, Y.; Ng, S.W.; Cate, H.S.; Sah, P.; Windels, F.; Kilpatrick, T.J.; Merson, T.D. Targeted Ablation of Oligodendrocytes Induces Axonal Pathology Independent of Overt Demyelination. J. Neurosci. 2012, 32, 8317–8330. [Google Scholar] [CrossRef] [PubMed]
  51. Saab, A.S.; Nave, K.-A. Myelin Dynamics: Protecting and Shaping Neuronal Functions. Curr. Opin. Neurobiol. 2017, 47, 104–112. [Google Scholar] [CrossRef]
  52. Snaidero, N.; Simons, M. Myelination at a Glance. J. Cell Sci. 2014, 127, 2999–3004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Xiao, L.; Ohayon, D.; McKenzie, I.A.; Sinclair-Wilson, A.; Wright, J.L.; Fudge, A.D.; Emery, B.; Li, H.; Richardson, W.D. Rapid Production of New Oligodendrocytes Is Required in the Earliest Stages of Motor-Skill Learning. Nat. Neurosci. 2016, 19, 1210–1217. [Google Scholar] [CrossRef] [PubMed]
  54. Gallo, V.; Zhou, J.; McBain, C.; Wright, P.; Knutson, P.; Armstrong, R. Oligodendrocyte Progenitor Cell Proliferation and Lineage Progression Are Regulated by Glutamate Receptor-Mediated K+ Channel Block. J. Neurosci. 1996, 16, 2659–2670. [Google Scholar] [CrossRef]
  55. Nagy, B.; Hovhannisyan, A.; Barzan, R.; Chen, T.-J.; Kukley, M. Different Patterns of Neuronal Activity Trigger Distinct Responses of Oligodendrocyte Precursor Cells in the Corpus Callosum. PLoS Biol. 2017, 15, e2001993. [Google Scholar] [CrossRef] [PubMed]
  56. Wake, H.; Lee, P.R.; Fields, R.D. Control of Local Protein Synthesis and Initial Events in Myelination by Action Potentials. Science 2011, 333, 1647–1651. [Google Scholar] [CrossRef] [Green Version]
  57. Yuan, X.; Eisen, A.M.; McBain, C.J.; Gallo, V. A Role for Glutamate and Its Receptors in the Regulation of Oligodendrocyte Development in Cerebellar Tissue Slices. Development 1998, 125, 2901–2914. [Google Scholar]
  58. Stevens, B.; Porta, S.; Haak, L.L.; Gallo, V.; Fields, R.D. Adenosine: A Neuron-Glial Transmitter Promoting Myelination in the CNS in Response to Action Potentials. Neuron 2002, 36, 855–868. [Google Scholar] [CrossRef] [Green Version]
  59. Ronzano, R.; Thetiot, M.; Lubetzki, C.; Desmazieres, A. Myelin Plasticity and Repair: Neuro-Glial Choir Sets the Tuning. Front. Cell. Neurosci. 2020, 14, 42. [Google Scholar] [CrossRef] [Green Version]
  60. Baraban, M.; Koudelka, S.; Lyons, D.A. Ca2+ Activity Signatures of Myelin Sheath Formation and Growth in Vivo. Nat. Neurosci. 2018, 21, 19–23. [Google Scholar] [CrossRef]
  61. Ortiz, F.C.; Habermacher, C.; Graciarena, M.; Houry, P.-Y.; Nishiyama, A.; Oumesmar, B.N.; Angulo, M.C. Neuronal Activity in Vivo Enhances Functional Myelin Repair. JCI Insight 2019, 4, e123434. [Google Scholar] [CrossRef] [Green Version]
  62. Trapp, B.D.; Nishiyama, A.; Cheng, D.; Macklin, W. Differentiation and Death of Premyelinating Oligodendrocytes in Developing Rodent Brain. J. Cell Biol. 1997, 137, 459–468. [Google Scholar] [CrossRef]
  63. Wilkins, A.; Majed, H.; Layfield, R.; Compston, A.; Chandran, S. Oligodendrocytes Promote Neuronal Survival and Axonal Length by Distinct Intracellular Mechanisms: A Novel Role for Oligodendrocyte-Derived Glial Cell Line-Derived Neurotrophic Factor. J. Neurosci. 2003, 23, 4967–4974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Simons, M.; Nave, K.-A. Oligodendrocytes: Myelination and Axonal Support. Cold Spring Harb. Perspect. Biol. 2016, 8, a020479. [Google Scholar] [CrossRef] [PubMed]
  65. Saab, A.S.; Tzvetanova, I.D.; Nave, K.-A. The Role of Myelin and Oligodendrocytes in Axonal Energy Metabolism. Curr. Opin. Neurobiol. 2013, 23, 1065–1072. [Google Scholar] [CrossRef] [PubMed]
  66. Meyer, N.; Richter, N.; Fan, Z.; Siemonsmeier, G.; Pivneva, T.; Jordan, P.; Steinhäuser, C.; Semtner, M.; Nolte, C.; Kettenmann, H. Oligodendrocytes in the Mouse Corpus Callosum Maintain Axonal Function by Delivery of Glucose. Cell Rep. 2018, 22, 2383–2394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Moore, S.; Meschkat, M.; Ruhwedel, T.; Trevisiol, A.; Tzvetanova, I.D.; Battefeld, A.; Kusch, K.; Kole, M.H.P.; Strenzke, N.; Möbius, W.; et al. A Role of Oligodendrocytes in Information Processing. Nat. Commun. 2020, 11, 5497. [Google Scholar] [CrossRef]
  68. Montag, D.; Giese, K.P.; Bartsch, U.; Martini, R.; Lang, Y.; Blüthmann, H.; Karthigasan, J.; Kirschner, D.A.; Wintergerst, E.S.; Nave, K.-A.; et al. Mice Deficient for the Glycoprotein Show Subtle Abnormalities in Myelin. Neuron 1994, 13, 229–246. [Google Scholar] [CrossRef]
  69. Griffiths, I. Axonal Swellings and Degeneration in Mice Lacking the Major Proteolipid of Myelin. Science 1998, 280, 1610–1613. [Google Scholar] [CrossRef]
  70. Garbern, J.Y.; Yool, D.A.; Moore, G.J.; Wilds, I.B.; Faulk, M.W.; Klugmann, M.; Nave, K.-A.; Sistermans, E.A.; van der Knaap, M.S.; Bird, T.D.; et al. Patients Lacking the Major CNS Myelin Protein, Proteolipid Protein 1, Develop Length-Dependent Axonal Degeneration in the Absence of Demyelination and Inflammation. Brain 2002, 125, 551–561. [Google Scholar] [CrossRef] [PubMed]
  71. Delaney, K.H.; Kwiecien, J.M.; Wegiel, J.; Wisniewski, H.M.; Percy, D.H.; Fletch, A.L. Familial Dysmyelination in a Long Evans Rat Mutant. Lab. Anim. Sci. 1995, 45, 547–553. [Google Scholar]
  72. Rosenbluth, J. Central Myelin in the Mouse Mutant Shiverer. J. Comp. Neurol. 1980, 194, 639–648. [Google Scholar] [CrossRef]
  73. Smith, C.M.; Cooksey, E.; Duncan, I.D. Myelin Loss Does Not Lead to Axonal Degeneration in a Long-Lived Model of Chronic Demyelination. J. Neurosci. 2013, 33, 2718–2727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Charcot, J.M. Histologie de La Sclerose En Plaques. Gaz. Hop. 1868, 41, 554–566. [Google Scholar]
  75. Zalc, B. One Hundred and Fifty Years Ago Charcot Reported Multiple Sclerosis as a New Neurological Disease. Brain 2018, 141, 3482–3488. [Google Scholar] [CrossRef] [PubMed]
  76. Thompson, A.J.; Banwell, B.L.; Barkhof, F.; Carroll, W.M.; Coetzee, T.; Comi, G.; Correale, J.; Fazekas, F.; Filippi, M.; Freedman, M.S.; et al. Diagnosis of Multiple Sclerosis: 2017 Revisions of the McDonald Criteria. Lancet Neurol. 2018, 17, 162–173. [Google Scholar] [CrossRef]
  77. Kuhlmann, T.; Ludwin, S.; Prat, A.; Antel, J.; Brück, W.; Lassmann, H. An Updated Histological Classification System for Multiple Sclerosis Lesions. Acta Neuropathol. 2017, 133, 13–24. [Google Scholar] [CrossRef]
  78. Lucchinetti, C.; Bruck, W.; Parisi, J.; Scheithauer, B.; Rodriguez, M.; Lassmann, H. Heterogeneity of Multiple Sclerosis Lesions: Implications for the Pathogenesis of Demyelination. Ann. Neurol. 2000, 47, 707–717. [Google Scholar] [CrossRef]
  79. Bramow, S.; Frischer, J.M.; Lassmann, H.; Koch-Henriksen, N.; Lucchinetti, C.F.; Sørensen, P.S.; Laursen, H. Demyelination versus Remyelination in Progressive Multiple Sclerosis. Brain 2010, 133, 2983–2998. [Google Scholar] [CrossRef] [Green Version]
  80. Coman, I.; Aigrot, M.S.; Seilhean, D.; Reynolds, R.; Girault, J.A.; Zalc, B.; Lubetzki, C. Nodal, Paranodal and Juxtaparanodal Axonal Proteins during Demyelination and Remyelination in Multiple Sclerosis. Brain 2006, 129, 3186–3195. [Google Scholar] [CrossRef] [PubMed]
  81. Craner, M.J.; Newcombe, J.; Black, J.A.; Hartle, C.; Cuzner, M.L.; Waxman, S.G. Molecular Changes in Neurons in Multiple Sclerosis: Altered Axonal Expression of Nav1.2 and Nav1.6 Sodium Channels and Na+/Ca2+ Exchanger. Proc. Natl. Acad. Sci. USA 2004, 101, 8168–8173. [Google Scholar] [CrossRef] [Green Version]
  82. Howell, O.W.; Palser, A.; Polito, A.; Melrose, S.; Zonta, B.; Scheiermann, C.; Vora, A.J.; Brophy, P.J.; Reynolds, R. Disruption of Neurofascin Localization Reveals Early Changes Preceding Demyelination and Remyelination in Multiple Sclerosis. Brain 2006, 129, 3173–3185. [Google Scholar] [CrossRef]
  83. Ronzano, R.; Roux, T.; Thetiot, M.; Aigrot, M.S.; Richard, L.; Lejeune, F.X.; Mazuir, E.; Vallat, J.M.; Lubetzki, C.; Desmazières, A. Microglia-Neuron Communication at Nodes of Ranvier Depends on Neuronal Activity through Potassium Release and Contributes to Myelin Repair. bioRxiv 2020. [Google Scholar] [CrossRef]
  84. Nikić, I.; Merkler, D.; Sorbara, C.; Brinkoetter, M.; Kreutzfeldt, M.; Bareyre, F.M.; Brück, W.; Bishop, D.; Misgeld, T.; Kerschensteiner, M. A Reversible Form of Axon Damage in Experimental Autoimmune Encephalomyelitis and Multiple Sclerosis. Nat. Med. 2011, 17, 495–499. [Google Scholar] [CrossRef]
  85. Lubetzki, C.; Zalc, B. Remyelination: A Brief History. Lancet Neurol. 2020, 19, 649. [Google Scholar] [CrossRef]
  86. Goldschmidt, T.; Antel, J.; König, F.B.; Brück, W.; Kuhlmann, T. Remyelination Capacity of the MS Brain Decreases with Disease Chronicity. Neurology 2009, 72, 1914–1921. [Google Scholar] [CrossRef] [PubMed]
  87. Patani, R.; Balaratnam, M.; Vora, A.; Reynolds, R. Remyelination Can Be Extensive in Multiple Sclerosis despite a Long Disease Course. Neuropathol. Appl. Neurobiol. 2007, 33, 277–287. [Google Scholar] [CrossRef] [PubMed]
  88. Prineas, J.W.; Barnard, R.O.; Revesz, T.; Kwon, E.E.; Sharer, L.; Cho, E.S. Multiple Sclerosis. Pathology of Recurrent Lesions. Brain 1993, 116 Pt 3, 681–693. [Google Scholar] [CrossRef]
  89. Franklin, R.J.M.; Frisén, J.; Lyons, D.A. Revisiting Remyelination: Towards a Consensus on the Regeneration of CNS Myelin. Semin. Cell Dev. Biol. 2020, S1084952120301579. [Google Scholar] [CrossRef]
  90. Levine, J.M.; Reynolds, R. Activation and Proliferation of Endogenous Oligodendrocyte Precursor Cells during Ethidium Bromide-Induced Demyelination. Exp. Neurol. 1999, 160, 333–347. [Google Scholar] [CrossRef]
  91. Fancy, S.P.J.; Zhao, C.; Franklin, R.J.M. Increased Expression of Nkx2.2 and Olig2 Identifies Reactive Oligodendrocyte Progenitor Cells Responding to Demyelination in the Adult CNS. Mol. Cell. Neurosci. 2004, 27, 247–254. [Google Scholar] [CrossRef] [PubMed]
  92. Miron, V.E.; Kuhlmann, T.; Antel, J.P. Cells of the Oligodendroglial Lineage, Myelination, and Remyelination. Biochim. Biophys. Acta BBA Mol. Basis Dis. 2011, 1812, 184–193. [Google Scholar] [CrossRef] [Green Version]
  93. Snaidero, N.; Schifferer, M.; Mezydlo, A.; Zalc, B.; Kerschensteiner, M.; Misgeld, T. Myelin Replacement Triggered by Single-Cell Demyelination in Mouse Cortex. Nat. Commun. 2020, 11, 4901. [Google Scholar] [CrossRef]
  94. Neely, S.A.; Williamson, J.M.; Klingseisen, A.; Zoupi, L.; Early, J.J.; Williams, A.; Lyons, D.A. New Oligodendrocytes Exhibit More Abundant and Accurate Myelin Regeneration than Those That Survive Demyelination. bioRxiv 2020. [Google Scholar] [CrossRef]
  95. Nait-Oumesmar, B.; Picard-Riera, N.; Kerninon, C.; Decker, L.; Seilhean, D.; Hoglinger, G.U.; Hirsch, E.C.; Reynolds, R.; Baron-Van Evercooren, A. Activation of the Subventricular Zone in Multiple Sclerosis: Evidence for Early Glial Progenitors. Proc. Natl. Acad. Sci. USA 2007, 104, 4694–4699. [Google Scholar] [CrossRef] [Green Version]
  96. Remaud, S.; Ortiz, F.C.; Perret-Jeanneret, M.; Aigrot, M.-S.; Gothié, J.-D.; Fekete, C.; Kvárta-Papp, Z.; Gereben, B.; Langui, D.; Lubetzki, C.; et al. Transient Hypothyroidism Favors Oligodendrocyte Generation Providing Functional Remyelination in the Adult Mouse Brain. eLife 2017, 6, e29996. [Google Scholar] [CrossRef]
  97. Xing, Y.L.; Röth, P.T.; Stratton, J.A.S.; Chuang, B.H.A.; Danne, J.; Ellis, S.L.; Ng, S.W.; Kilpatrick, T.J.; Merson, T.D. Adult Neural Precursor Cells from the Subventricular Zone Contribute Significantly to Oligodendrocyte Regeneration and Remyelination. J. Neurosci. 2014, 34, 14128–14146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Lubetzki, C.; Gansmüller, A.; Lachapelle, F.; Lombrail, P.; Gumpel, M. Myelination by Oligodendrocytes Isolated from 4-6-Week-Old Rat Central Nervous System and Transplanted into Newborn Shiverer Brain. J. Neurol. Sci. 1988, 88, 161–175. [Google Scholar] [CrossRef]
  99. Yeung, M.S.Y.; Zdunek, S.; Bergmann, O.; Bernard, S.; Salehpour, M.; Alkass, K.; Perl, S.; Tisdale, J.; Possnert, G.; Brundin, L.; et al. Dynamics of Oligodendrocyte Generation and Myelination in the Human Brain. Cell 2014, 159, 766–774. [Google Scholar] [CrossRef] [Green Version]
  100. Yeung, M.S.Y.; Djelloul, M.; Steiner, E.; Bernard, S.; Salehpour, M.; Possnert, G.; Brundin, L.; Frisén, J. Dynamics of Oligodendrocyte Generation in Multiple Sclerosis. Nature 2019, 566, 538–542. [Google Scholar] [CrossRef]
  101. Lucchinetti, C.; Brück, W.; Parisi, J.; Scheithauer, B.; Rodriguez, M.; Lassmann, H. A Quantitative Analysis of Oligodendrocytes in Multiple Sclerosis Lesions. A Study of 113 Cases. Brain 1999, 122 Pt 12, 2279–2295. [Google Scholar] [CrossRef]
  102. Brown, R.A.; Narayanan, S.; Arnold, D.L. Imaging of Repeated Episodes of Demyelination and Remyelination in Multiple Sclerosis. Neuroimage Clin. 2014, 6, 20–25. [Google Scholar] [CrossRef] [Green Version]
  103. Franklin, R.J.M. Why Does Remyelination Fail in Multiple Sclerosis? Nat. Rev. Neurosci. 2002, 3, 705–714. [Google Scholar] [CrossRef]
  104. Ruckh, J.M.; Zhao, J.-W.; Shadrach, J.L.; van Wijngaarden, P.; Rao, T.N.; Wagers, A.J.; Franklin, R.J.M. Rejuvenation of Regeneration in the Aging Central Nervous System. Cell Stem Cell 2012, 10, 96–103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Sim, F.J.; Zhao, C.; Penderis, J.; Franklin, R.J.M. The Age-Related Decrease in CNS Remyelination Efficiency Is Attributable to an Impairment of Both Oligodendrocyte Progenitor Recruitment and Differentiation. J. Neurosci. 2002, 22, 2451–2459. [Google Scholar] [CrossRef] [PubMed]
  106. Absinta, M.; Sati, P.; Schindler, M.; Leibovitch, E.C.; Ohayon, J.; Wu, T.; Meani, A.; Filippi, M.; Jacobson, S.; Cortese, I.C.M.; et al. Persistent 7-Tesla Phase Rim Predicts Poor Outcome in New Multiple Sclerosis Patient Lesions. J. Clin. Investig. 2016, 126, 2597–2609. [Google Scholar] [CrossRef]
  107. Dal-Bianco, A.; Grabner, G.; Kronnerwetter, C.; Weber, M.; Höftberger, R.; Berger, T.; Auff, E.; Leutmezer, F.; Trattnig, S.; Lassmann, H.; et al. Slow Expansion of Multiple Sclerosis Iron Rim Lesions: Pathology and 7 T Magnetic Resonance Imaging. Acta Neuropathol. 2017, 133, 25–42. [Google Scholar] [CrossRef] [Green Version]
  108. Luchetti, S.; Fransen, N.L.; van Eden, C.G.; Ramaglia, V.; Mason, M.; Huitinga, I. Progressive Multiple Sclerosis Patients Show Substantial Lesion Activity That Correlates with Clinical Disease Severity and Sex: A Retrospective Autopsy Cohort Analysis. Acta Neuropathol. 2018, 135, 511–528. [Google Scholar] [CrossRef] [Green Version]
  109. Kornek, B.; Lassmann, H. Neuropathology of Multiple Sclerosis—New Concepts. Brain Res. Bull. 2003, 61, 321–326. [Google Scholar] [CrossRef]
  110. Mei, F.; Lehmann-Horn, K.; Shen, Y.-A.A.; Rankin, K.A.; Stebbins, K.J.; Lorrain, D.S.; Pekarek, K.; Sagan, S.A.; Xiao, L.; Teuscher, C.; et al. Accelerated Remyelination during Inflammatory Demyelination Prevents Axonal Loss and Improves Functional Recovery. eLife 2019, 5, e18246. [Google Scholar] [CrossRef]
  111. Piaton, G.; Aigrot, M.-S.; Williams, A.; Moyon, S.; Tepavcevic, V.; Moutkine, I.; Gras, J.; Matho, K.S.; Schmitt, A.; Soellner, H.; et al. Class 3 Semaphorins Influence Oligodendrocyte Precursor Recruitment and Remyelination in Adult Central Nervous System. Brain 2011, 134, 1156–1167. [Google Scholar] [CrossRef] [Green Version]
  112. Tepavčević, V.; Kerninon, C.; Aigrot, M.S.; Meppiel, E.; Mozafari, S.; Arnould-Laurent, R.; Ravassard, P.; Kennedy, T.E.; Nait-Oumesmar, B.; Lubetzki, C. Early Netrin-1 Expression Impairs Central Nervous System Remyelination: Netrin-1 and CNS Remyelination. Ann. Neurol. 2014, 76, 252–268. [Google Scholar] [CrossRef]
  113. Ishibashi, T.; Dakin, K.A.; Stevens, B.; Lee, P.R.; Kozlov, S.V.; Stewart, C.L.; Fields, R.D. Astrocytes Promote Myelination in Response to Electrical Impulses. Neuron 2006, 49, 823–832. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Stankoff, B.; Aigrot, M.-S.; Noël, F.; Wattilliaux, A.; Zalc, B.; Lubetzki, C. Ciliary Neurotrophic Factor (CNTF) Enhances Myelin Formation: A Novel Role for CNTF and CNTF-Related Molecules. J. Neurosci. 2002, 22, 9221–9227. [Google Scholar] [CrossRef] [PubMed]
  115. Craner, M.J. Abnormal Sodium Channel Distribution in Optic Nerve Axons in a Model of Inflammatory Demyelination. Brain 2003, 126, 1552–1561. [Google Scholar] [CrossRef] [PubMed]
  116. Orthmann-Murphy, J.; Call, C.L.; Molina-Castro, G.C.; Hsieh, Y.C.; Rasband, M.N.; Calabresi, P.A.; Bergles, D.E. Remyelination Alters the Pattern of Myelin in the Cerebral Cortex. eLife 2020, 9, e56621. [Google Scholar] [CrossRef] [PubMed]
  117. Freeman, S.A.; Desmazières, A.; Fricker, D.; Lubetzki, C.; Sol-Foulon, N. Mechanisms of Sodium Channel Clustering and Its Influence on Axonal Impulse Conduction. Cell. Mol. Life Sci. 2016, 73, 723–735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Kaplan, M.R.; Cho, M.-H.; Ullian, E.M.; Isom, L.L.; Levinson, S.R.; Barres, B.A. Differential Control of Clustering of the Sodium Channels Nav1.2 and Nav1.6 at Developing CNS Nodes of Ranvier. Neuron 2001, 30, 105–119. [Google Scholar] [CrossRef] [Green Version]
  119. John, G.R.; Shankar, S.L.; Shafit-Zagardo, B.; Massimi, A.; Lee, S.C.; Raine, C.S.; Brosnan, C.F. Multiple Sclerosis: Re-Expression of a Developmental Pathway That Restricts Oligodendrocyte Maturation. Nat. Med. 2002, 8, 1115–1121. [Google Scholar] [CrossRef] [PubMed]
  120. Lau, L.W.; Cua, R.; Keough, M.B.; Haylock-Jacobs, S.; Yong, V.W. Pathophysiology of the Brain Extracellular Matrix: A New Target for Remyelination. Nat. Rev. Neurosci. 2013, 14, 722–729. [Google Scholar] [CrossRef] [PubMed]
  121. Kuboyama, K.; Tanga, N.; Suzuki, R.; Fujikawa, A.; Noda, M. Protamine Neutralizes Chondroitin Sulfate Proteoglycan-Mediated Inhibition of Oligodendrocyte Differentiation. PLoS ONE 2017, 12, e0189164. [Google Scholar] [CrossRef]
  122. Luo, F.; Tran, A.P.; Xin, L.; Sanapala, C.; Lang, B.T.; Silver, J.; Yang, Y. Modulation of Proteoglycan Receptor PTPσ Enhances MMP-2 Activity to Promote Recovery from Multiple Sclerosis. Nat. Commun. 2018, 9, 4126. [Google Scholar] [CrossRef]
  123. Pekny, M.; Wilhelmsson, U.; Pekna, M. The Dual Role of Astrocyte Activation and Reactive Gliosis. Neurosci. Lett. 2014, 565, 30–38. [Google Scholar] [CrossRef]
  124. Franklin, R.J.; Crang, A.J.; Blakemore, W.F. Transplanted Type-1 Astrocytes Facilitate Repair of Demyelinating Lesions by Host Oligodendrocytes in Adult Rat Spinal Cord. J. Neurocytol. 1991, 20, 420–430. [Google Scholar] [CrossRef]
  125. Neumann, H.; Kotter, M.R.; Franklin, R.J.M. Debris Clearance by Microglia: An Essential Link between Degeneration and Regeneration. Brain 2008, 132, 288–295. [Google Scholar] [CrossRef]
  126. Kotter, M.R.; Zhao, C.; van Rooijen, N.; Franklin, R.J.M. Macrophage-Depletion Induced Impairment of Experimental CNS Remyelination Is Associated with a Reduced Oligodendrocyte Progenitor Cell Response and Altered Growth Factor Expression. Neurobiol. Dis. 2005, 18, 166–175. [Google Scholar] [CrossRef]
  127. Bauer, J.; Sminia, T.; Wouterlood, F.G.; Dijkstra, C.D. Phagocytic Activity of Macrophages and Microglial Cells during the Course of Acute and Chronic Relapsing Experimental Autoimmune Encephalomyelitis. J. Neurosci. Res. 1994, 38, 365–375. [Google Scholar] [CrossRef] [PubMed]
  128. Lloyd, A.F.; Davies, C.L.; Miron, V.E. Microglia: Origins, Homeostasis, and Roles in Myelin Repair. Curr. Opin. Neurobiol. 2017, 47, 113–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Lloyd, A.F.; Miron, V.E. The Pro-Remyelination Properties of Microglia in the Central Nervous System. Nat. Rev. Neurol. 2019, 15, 447–458. [Google Scholar] [CrossRef] [PubMed]
  130. Miron, V.E.; Boyd, A.; Zhao, J.-W.; Yuen, T.J.; Ruckh, J.M.; Shadrach, J.L.; van Wijngaarden, P.; Wagers, A.J.; Williams, A.; Franklin, R.J.M.; et al. M2 Microglia and Macrophages Drive Oligodendrocyte Differentiation during CNS Remyelination. Nat. Neurosci. 2013, 16, 1211–1218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Miron, V.E. Microglia-Driven Regulation of Oligodendrocyte Lineage Cells, Myelination, and Remyelination. J. Leukoc. Biol. 2017, 101, 1103–1108. [Google Scholar] [CrossRef] [PubMed]
  132. Dziembowska, M.; Tham, T.N.; Lau, P.; Vitry, S.; Lazarini, F.; Dubois-Dalcq, M. A Role for CXCR4 Signaling in Survival and Migration of Neural and Oligodendrocyte Precursors. Glia 2005, 50, 258–269. [Google Scholar] [CrossRef]
  133. Connor, J.R.; Menzies, S.L. Relationship of Iron to Oligodendrocytes and Myelination. Glia 1996, 17, 83–93. [Google Scholar] [CrossRef]
  134. Shields, S.A.; Gilson, J.M.; Blakemore, W.F.; Franklin, R.J. Remyelination Occurs as Extensively but More Slowly in Old Rats Compared to Young Rats Following Gliotoxin-Induced CNS Demyelination. Glia 1999, 28, 77–83. [Google Scholar] [CrossRef]
  135. Wolswijk, G. Chronic Stage Multiple Sclerosis Lesions Contain a Relatively Quiescent Population of Oligodendrocyte Precursor Cells. J. Neurosci. 1998, 18, 601–609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Giovannoni, G.; Bermel, R.; Phillips, T.; Rudick, R. A Brief History of NEDA. Mult. Scler. Relat. Disord. 2018, 20, 228–230. [Google Scholar] [CrossRef]
  137. Friese, M.A.; Schattling, B.; Fugger, L. Mechanisms of Neurodegeneration and Axonal Dysfunction in Multiple Sclerosis. Nat. Rev. Neurol. 2014, 10, 225–238. [Google Scholar] [CrossRef] [PubMed]
  138. Singh, S.; Dallenga, T.; Winkler, A.; Roemer, S.; Maruschak, B.; Siebert, H.; Brück, W.; Stadelmann, C. Relationship of Acute Axonal Damage, Wallerian Degeneration, and Clinical Disability in Multiple Sclerosis. J. Neuroinflamm. 2017, 14, 57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Hooijmans, C.R.; Hlavica, M.; Schuler, F.A.F.; Good, N.; Good, A.; Baumgartner, L.; Galeno, G.; Schneider, M.P.; Jung, T.; de Vries, R.; et al. Remyelination Promoting Therapies in Multiple Sclerosis Animal Models: A Systematic Review and Meta-Analysis. Sci. Rep. 2019, 9, 822. [Google Scholar] [CrossRef] [Green Version]
  140. Bodini, B.; Stankoff, B. Imaging Central Nervous System Demyelination and Remyelination by Positron-Emission Tomography. Brain Plast. 2016, 2, 93–98. [Google Scholar] [CrossRef] [Green Version]
  141. Mei, F.; Fancy, S.P.J.; Shen, Y.-A.A.; Niu, J.; Zhao, C.; Presley, B.; Miao, E.; Lee, S.; Mayoral, S.R.; Redmond, S.A.; et al. Micropillar Arrays as a High-Throughput Screening Platform for Therapeutics in Multiple Sclerosis. Nat. Med. 2014, 20, 954–960. [Google Scholar] [CrossRef] [Green Version]
  142. Li, Z.; He, Y.; Fan, S.; Sun, B. Clemastine Rescues Behavioral Changes and Enhances Remyelination in the Cuprizone Mouse Model of Demyelination. Neurosci. Bull. 2015, 31, 617–625. [Google Scholar] [CrossRef] [PubMed]
  143. Green, A.J.; Gelfand, J.M.; Cree, B.A.; Bevan, C.; Boscardin, W.J.; Mei, F.; Inman, J.; Arnow, S.; Devereux, M.; Abounasr, A.; et al. Clemastine Fumarate as a Remyelinating Therapy for Multiple Sclerosis (ReBUILD): A Randomised, Controlled, Double-Blind, Crossover Trial. Lancet 2017, 390, 2481–2489. [Google Scholar] [CrossRef] [Green Version]
  144. Schwartzbach, C.J.; Grove, R.A.; Brown, R.; Tompson, D.; Then Bergh, F.; Arnold, D.L. Lesion Remyelinating Activity of GSK239512 versus Placebo in Patients with Relapsing-Remitting Multiple Sclerosis: A Randomised, Single-Blind, Phase II Study. J. Neurol. 2017, 264, 304–315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Deshmukh, V.A.; Tardif, V.; Lyssiotis, C.A.; Green, C.C.; Kerman, B.; Kim, H.J.; Padmanabhan, K.; Swoboda, J.G.; Ahmad, I.; Kondo, T.; et al. A Regenerative Approach to the Treatment of Multiple Sclerosis. Nature 2013, 502, 327–332. [Google Scholar] [CrossRef] [Green Version]
  146. Bongarzone, E.R.; Howard, S.G.; Schonmann, V.; Campagnoni, A.T. Identification of the Dopamine D3 Receptor in Oligodendrocyte Precursors: Potential Role in Regulating Differentiation and Myelin Formation. J. Neurosci. 1998, 18, 5344–5353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Zhang, Y.; Xu, H.; Jiang, W.; Xiao, L.; Yan, B.; He, J.; Wang, Y.; Bi, X.; Li, X.; Kong, J. Quetiapine Alleviates the Cuprizone-Induced White Matter Pathology in the Brain of C57BL/6 Mouse. Schizophr. Res. 2008, 106, 182–191. [Google Scholar] [CrossRef]
  148. Gonzalez, G.A.; Hofer, M.P.; Syed, Y.A.; Amaral, A.I.; Rundle, J.; Rahman, S.; Zhao, C.; Kotter, M.R.N. Tamoxifen Accelerates the Repair of Demyelinated Lesions in the Central Nervous System. Sci. Rep. 2016, 6, 31599. [Google Scholar] [CrossRef]
  149. Chan, J.R.; Phillips, L.J.; Glaser, M. Glucocorticoids and Progestins Signal the Initiation and Enhance the Rate of Myelin Formation. Proc. Natl. Acad. Sci. USA 1998, 95, 10459–10464. [Google Scholar] [CrossRef] [Green Version]
  150. Najm, F.J.; Madhavan, M.; Zaremba, A.; Shick, E.; Karl, R.T.; Factor, D.C.; Miller, T.E.; Nevin, Z.S.; Kantor, C.; Sargent, A.; et al. Drug-Based Modulation of Endogenous Stem Cells Promotes Functional Remyelination in Vivo. Nature 2015, 522, 216–220. [Google Scholar] [CrossRef] [Green Version]
  151. Neumann, B.; Baror, R.; Zhao, C.; Segel, M.; Dietmann, S.; Rawji, K.S.; Foerster, S.; McClain, C.R.; Chalut, K.; van Wijngaarden, P.; et al. Metformin Restores CNS Remyelination Capacity by Rejuvenating Aged Stem Cells. Cell Stem Cell 2019, 25, 473–485.e8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Göttle, P.; Manousi, A.; Kremer, D.; Reiche, L.; Hartung, H.-P.; Küry, P. Teriflunomide Promotes Oligodendroglial Differentiation and Myelination. J. Neuroinflamm. 2018, 15, 76. [Google Scholar] [CrossRef] [Green Version]
  153. Mannioui, A.; Vauzanges, Q.; Fini, J.B.; Henriet, E.; Sekizar, S.; Azoyan, L.; Thomas, J.L.; Pasquier, D.D.; Giovannangeli, C.; Demeneix, B.; et al. The Xenopus Tadpole: An in Vivo Model to Screen Drugs Favoring Remyelination. Mult. Scler. 2018, 24, 1421–1432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Miron, V.E.; Rajasekharan, S.; Jarjour, A.A.; Zamvil, S.S.; Kennedy, T.E.; Antel, J.P. Simvastatin Regulates Oligodendroglial Process Dynamics and Survival. Glia 2007, 55, 130–143. [Google Scholar] [CrossRef]
  155. Chataway, J.; Schuerer, N.; Alsanousi, A.; Chan, D.; MacManus, D.; Hunter, K.; Anderson, V.; Bangham, C.R.M.; Clegg, S.; Nielsen, C.; et al. Effect of High-Dose Simvastatin on Brain Atrophy and Disability in Secondary Progressive Multiple Sclerosis (MS-STAT): A Randomised, Placebo-Controlled, Phase 2 Trial. Lancet 2014, 383, 2213–2221. [Google Scholar] [CrossRef] [Green Version]
  156. Sorensen, P.S.; Lycke, J.; Erälinna, J.-P.; Edland, A.; Wu, X.; Frederiksen, J.L.; Oturai, A.; Malmeström, C.; Stenager, E.; Sellebjerg, F.; et al. Simvastatin as Add-on Therapy to Interferon Beta-1a for Relapsing-Remitting Multiple Sclerosis (SIMCOMBIN Study): A Placebo-Controlled Randomised Phase 4 Trial. Lancet Neurol. 2011, 10, 691–701. [Google Scholar] [CrossRef]
  157. Shen, S.; Sandoval, J.; Swiss, V.A.; Li, J.; Dupree, J.; Franklin, R.J.M.; Casaccia-Bonnefil, P. Age-Dependent Epigenetic Control of Differentiation Inhibitors Is Critical for Remyelination Efficiency. Nat. Neurosci. 2008, 11, 1024–1034. [Google Scholar] [CrossRef] [Green Version]
  158. Moreau-Fauvarque, C.; Kumanogoh, A.; Camand, E.; Jaillard, C.; Barbin, G.; Boquet, I.; Love, C.; Jones, E.Y.; Kikutani, H.; Lubetzki, C.; et al. The Transmembrane Semaphorin Sema4D/CD100, an Inhibitor of Axonal Growth, Is Expressed on Oligodendrocytes and Upregulated after CNS Lesion. J. Neurosci. 2003, 23, 9229–9239. [Google Scholar] [CrossRef] [PubMed]
  159. Smith, E.S.; Jonason, A.; Reilly, C.; Veeraraghavan, J.; Fisher, T.; Doherty, M.; Klimatcheva, E.; Mallow, C.; Cornelius, C.; Leonard, J.E.; et al. SEMA4D Compromises Blood–Brain Barrier, Activates Microglia, and Inhibits Remyelination in Neurodegenerative Disease. Neurobiol. Dis. 2015, 73, 254–268. [Google Scholar] [CrossRef] [Green Version]
  160. LaGanke, C.; Samkoff, L.; Edwards, K.; Jung Henson, L.; Repovic, P.; Lynch, S.; Stone, L.; Mattson, D.; Galluzzi, A.; Fisher, T.L.; et al. Safety/Tolerability of the Anti-Semaphorin 4D Antibody VX15/2503 in a Randomized Phase 1 Trial. Neurol. Neuroimmunol. Neuroinflamm. 2017, 4, e367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Derfuss, T.; Curtin, F.; Guebelin, C.; Bridel, C.; Rasenack, M.; Matthey, A.; Du Pasquier, R.; Schluep, M.; Desmeules, J.; Lang, A.B.; et al. A Phase IIa Randomised Clinical Study of GNbAC1, a Humanised Monoclonal Antibody against the Envelope Protein of Multiple Sclerosis-Associated Endogenous Retrovirus in Multiple Sclerosis Patients. Mult. Scler. 2015, 21, 885–893. [Google Scholar] [CrossRef] [Green Version]
  162. Mi, S.; Hu, B.; Hahm, K.; Luo, Y.; Kam Hui, E.S.; Yuan, Q.; Wong, W.M.; Wang, L.; Su, H.; Chu, T.-H.; et al. LINGO-1 Antagonist Promotes Spinal Cord Remyelination and Axonal Integrity in MOG-Induced Experimental Autoimmune Encephalomyelitis. Nat. Med. 2007, 13, 1228–1233. [Google Scholar] [CrossRef]
  163. Kremer, D.; Schichel, T.; Förster, M.; Tzekova, N.; Bernard, C.; van der Valk, P.; van Horssen, J.; Hartung, H.-P.; Perron, H.; Küry, P. Human Endogenous Retrovirus Type W Envelope Protein Inhibits Oligodendroglial Precursor Cell Differentiation: Retroviruses and Myelin Repair. Ann. Neurol. 2013, 74, 721–732. [Google Scholar] [CrossRef] [PubMed]
  164. Cadavid, D.; Balcer, L.; Galetta, S.; Aktas, O.; Ziemssen, T.; Vanopdenbosch, L.; Frederiksen, J.; Skeen, M.; Jaffe, G.J.; Butzkueven, H.; et al. Safety and Efficacy of Opicinumab in Acute Optic Neuritis (RENEW): A Randomised, Placebo-Controlled, Phase 2 Trial. Lancet Neurol. 2017, 16, 189–199. [Google Scholar] [CrossRef]
  165. Cadavid, D.; Mellion, M.; Hupperts, R.; Edwards, K.R.; Calabresi, P.A.; Drulović, J.; Giovannoni, G.; Hartung, H.-P.; Arnold, D.L.; Fisher, E.; et al. Safety and Efficacy of Opicinumab in Patients with Relapsing Multiple Sclerosis (SYNERGY): A Randomised, Placebo-Controlled, Phase 2 Trial. Lancet Neurol. 2019, 18, 845–856. [Google Scholar] [CrossRef]
  166. Martin, E.; Aigrot, M.-S.; Grenningloh, R.; Stankoff, B.; Lubetzki, C.; Boschert, U.; Zalc, B. Bruton’s Tyrosine Kinase Inhibition Promotes Myelin Repair. Brain Plast. 2020, 5, 123–133. [Google Scholar] [CrossRef]
  167. Montalban, X.; Arnold, D.L.; Weber, M.S.; Staikov, I.; Piasecka-Stryczynska, K.; Willmer, J.; Martin, E.C.; Dangond, F.; Syed, S.; Wolinsky, J.S. Placebo-Controlled Trial of an Oral BTK Inhibitor in Multiple Sclerosis. N. Engl. J. Med. 2019, 380, 2406–2417. [Google Scholar] [CrossRef]
  168. Chandraratna, R.A.; Noelle, R.J.; Nowak, E.C. Treatment with Retinoid X Receptor Agonist IRX4204 Ameliorates Experimental Autoimmune Encephalomyelitis. Am. J. Transl. Res. 2016, 8, 1016–1026. [Google Scholar] [PubMed]
  169. Ljunggren-Rose, Å.; Natarajan, C.; Matta, P.; Pandey, A.; Upender, I.; Sriram, S. Anacardic Acid Induces IL-33 and Promotes Remyelination in CNS. Proc. Natl. Acad. Sci. USA 2020, 117, 21527–21535. [Google Scholar] [CrossRef]
  170. Melero-Jerez, C.; Fernández-Gómez, B.; Lebrón-Galán, R.; Ortega, M.C.; Sánchez-de Lara, I.; Ojalvo, A.C.; Clemente, D.; Castro, F. Myeloid-derived Suppressor Cells Support Remyelination in a Murine Model of Multiple Sclerosis by Promoting Oligodendrocyte Precursor Cell Survival, Proliferation, and Differentiation. Glia 2020, 69, 905–924. [Google Scholar] [CrossRef]
  171. Matías-Guiu, J.; Matías-Guiu, J.A.; Montero-Escribano, P.; Barcia, J.A.; Canales-Aguirre, A.A.; Mateos-Diaz, J.C.; Gómez-Pinedo, U. Particles Containing Cells as a Strategy to Promote Remyelination in Patients With Multiple Sclerosis. Front. Neurol. 2020, 11, 638. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of myelin and axonal domain alteration and reformation in MS and its experimental models. Compared to normal condition (1), myelinated axon first undergo alteration at the nodal area in NAWM (normal appearing white matter) (2), with macrophage attack, paranodal axo-glial junction disruption and the appearance of reversible focal axonal damage (FAD). In demyelinating lesions (3), myelin degradation and nodal marker redistribution are associated with reduced conduction velocity or conduction blocks and associated functional deficits. These insults can lead to neurodegeneration with time (4), but an endogenous remyelination process exists (5), which is promoted by neuronal activity, and can lead to the restoration of myelinated fiber organization and efficient axonal conduction (6). AIS: axon initial segment.
Figure 1. Schematic representation of myelin and axonal domain alteration and reformation in MS and its experimental models. Compared to normal condition (1), myelinated axon first undergo alteration at the nodal area in NAWM (normal appearing white matter) (2), with macrophage attack, paranodal axo-glial junction disruption and the appearance of reversible focal axonal damage (FAD). In demyelinating lesions (3), myelin degradation and nodal marker redistribution are associated with reduced conduction velocity or conduction blocks and associated functional deficits. These insults can lead to neurodegeneration with time (4), but an endogenous remyelination process exists (5), which is promoted by neuronal activity, and can lead to the restoration of myelinated fiber organization and efficient axonal conduction (6). AIS: axon initial segment.
Life 11 00231 g001
Figure 2. Schematic representation of modulators of the remyelination process. Remyelination depends on the recruitment of OPCs at the demyelinated lesion and their differentiation into oligodendrocytes, which will remyelinate demyelinated axons. Neuronal activity promotes multiple steps of this pathway, including the myelination process per se (choice of the axonal target, myelin stabilization). Other key players correspond to microglia and astrocytes, which secrete various factors modulating the different steps of the (re)myelination process. OPC: oligodendrocyte precursor cell; OL: oligodendrocyte.AIS: Axon Initial Segment.
Figure 2. Schematic representation of modulators of the remyelination process. Remyelination depends on the recruitment of OPCs at the demyelinated lesion and their differentiation into oligodendrocytes, which will remyelinate demyelinated axons. Neuronal activity promotes multiple steps of this pathway, including the myelination process per se (choice of the axonal target, myelin stabilization). Other key players correspond to microglia and astrocytes, which secrete various factors modulating the different steps of the (re)myelination process. OPC: oligodendrocyte precursor cell; OL: oligodendrocyte.AIS: Axon Initial Segment.
Life 11 00231 g002
Table 1. Molecules tested in clinical trials for their remyelinating potential. AAR: Annualized Atrophy Rate in MRI; ARR: Annualized Relapse Rate; ACTH: Adrenocorticotropic Hormone; BTK: Bruton’s Tyrosine Kinase; CDP: 6 month Confirmed Disability Progression; EAE: Experimental Autoimmune Encephalomyelitis; EDSS: Expanded Disability Status Scale; HERV: Human Endogenous Retrovirus; IFN: Interferon Gamma; MTR: Magnetization Transfer Ratio; MWF: Myelin Water Fraction; LPC: Lysophosphatydilcholine-Induced Demyelination; mAb: Monoclonal Antibody; NfL: Neurofilament; ODRS: Overall Disability Response Score; OL: Oligodendrocyte; OPC: Oligodendrocyte Progenitor Cell; RNFL: Retinal Nerve Fiber layer; T25FW: Timed 25 Feet Walk Test; VEP: Visual Evoked Potential. Clinical trial data listed from ClinicalTrials.gov accessed on 7 March 2021.
Table 1. Molecules tested in clinical trials for their remyelinating potential. AAR: Annualized Atrophy Rate in MRI; ARR: Annualized Relapse Rate; ACTH: Adrenocorticotropic Hormone; BTK: Bruton’s Tyrosine Kinase; CDP: 6 month Confirmed Disability Progression; EAE: Experimental Autoimmune Encephalomyelitis; EDSS: Expanded Disability Status Scale; HERV: Human Endogenous Retrovirus; IFN: Interferon Gamma; MTR: Magnetization Transfer Ratio; MWF: Myelin Water Fraction; LPC: Lysophosphatydilcholine-Induced Demyelination; mAb: Monoclonal Antibody; NfL: Neurofilament; ODRS: Overall Disability Response Score; OL: Oligodendrocyte; OPC: Oligodendrocyte Progenitor Cell; RNFL: Retinal Nerve Fiber layer; T25FW: Timed 25 Feet Walk Test; VEP: Visual Evoked Potential. Clinical trial data listed from ClinicalTrials.gov accessed on 7 March 2021.
Drug NameMode of ActionPreclinical StudiesClinical Studies
ModelEffectStudyPopulationOutcomeResults
ACTHCorticotropin hormoneCell cultureOPC differentiation/OL maturationPhase IV (NCT02446886)RMSMWF over 12 monthsCompleted; results pending
Bexarotene/IRX4204Retinoid X receptor γ agonistEAEOPC differentiation/reduced EAE severityPhase IIa ISRCTN14265371RMSMTR; VEP latencyImprovement in MTR in a lesion subset; reduction in VEP latency ; serious side effects
Clemastine fumarateH1 and M1/M3 receptors antagonistMicropillar array screen; LPC; cuprizoneOPC differentiation/OL maturation/reduced EAE severityPhase II ReBUILD trial
Phase II Recover Trial
Chronic optic neuritis
Acute optic neuritis
P100 VEP latency
P100 VEP latency; RNFL thickness; EDSS; MWF
Reduced P100 VEP latency in chronic optic neuropathy without clinical improvement
Ongoing
DomperidoneD2/D3 dopamine receptor antagonistLPCOL maturationPhase II NCT02308137
Phase II NCT02493049
SPMS
RRMS
T25WF
MRI lesions; EDSS
Completed; results pending
Completed; results pending
GSK239512H3 receptor antagonistCuprizoneOPC differentiation/OL maturationPhase II
NCT01772199
RRMSMTR activitySmall improvement in lesion remyelination
OpicinumabmAb against LINGO 1EAE; LPCOPC differentiation/OL maturationPhase IIa RENEW trial
Phase IIb SYNERGY trial
Phase II AFFINITY trial
Acute optic neuritis
RMS
RMS
P100 VEP latency;
RNFL thickness; MRI lesions
EDSS progression
ODRS; MTR
Minor reduction in P100 VEP latency
No improvement vs. placebo
No improvement vs. placebo
QuetiapineAntipsychotic drugCuprizoneOL maturation; inhibition of activated microgliaPhase I/II NCT020087631RMSTolerance; EDSSCompleted; results pending
Temelimab/GNbAC1mAb against HERV envelope proteinIn vitro OPC culturesOPC differentiationPhase IIa CHANGE-MS trial
Phase IIb ANGEL-MS trial
RMS
RRMS
MRI lesions; morphometrySignificant reduction on cortical and thalamic atrophy; no decrease in new lesions
SAR442168
Evobrutinib/M2951
BTK inhibitorLPC demyelinized cultures1.7× improved remyelination comparing to placeboPhase III HERCULES trial
Phase III PERSEUS trial
2 PHASE III GEMINI1 and GEMINI2
Phase II NCT02975349
SPMS
PPMS
RMS
RMS
CDP;AAR; NfL
MRI, ARR, EDSS
Ongoing
Reduction in enhancing lesions and ARR, no effect on EDSS, transaminase elevation
SimvastatinHMG-CoA reductase inhibitorEAEOPC differentiation/prevent EAEPhase II
MS-STAT trial
Phase III MS-STAT2 trial
Phase IV SIMCOMBIN trial
SPMS
SPMS
RRMS
AAR
EDSS
Multiparameter clinical outcomes
ARR; MRI
43% reduction of brain atrophy
Ongoing
No additional benefit to IFN treatment
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pantazou, V.; Roux, T.; Oliveira Moreira, V.; Lubetzki, C.; Desmazières, A. Interaction between Neurons and the Oligodendroglial Lineage in Multiple Sclerosis and Its Preclinical Models. Life 2021, 11, 231. https://0-doi-org.brum.beds.ac.uk/10.3390/life11030231

AMA Style

Pantazou V, Roux T, Oliveira Moreira V, Lubetzki C, Desmazières A. Interaction between Neurons and the Oligodendroglial Lineage in Multiple Sclerosis and Its Preclinical Models. Life. 2021; 11(3):231. https://0-doi-org.brum.beds.ac.uk/10.3390/life11030231

Chicago/Turabian Style

Pantazou, Vasiliki, Thomas Roux, Vanessa Oliveira Moreira, Catherine Lubetzki, and Anne Desmazières. 2021. "Interaction between Neurons and the Oligodendroglial Lineage in Multiple Sclerosis and Its Preclinical Models" Life 11, no. 3: 231. https://0-doi-org.brum.beds.ac.uk/10.3390/life11030231

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop