Skip to main content
Neuroscience Bulletin logoLink to Neuroscience Bulletin
. 2024 Jul 23;40(10):1573–1589. doi: 10.1007/s12264-024-01261-8

A New Acquaintance of Oligodendrocyte Precursor Cells in the Central Nervous System

Zexuan Ma 1,#, Wei Zhang 1,#, Chenmeng Wang 1,2,#, Yixun Su 2, Chenju Yi 2,3,4,, Jianqin Niu 1,5,
PMCID: PMC11422404  PMID: 39042298

Abstract

Oligodendrocyte precursor cells (OPCs) are a heterogeneous multipotent population in the central nervous system (CNS) that appear during embryogenesis and persist as resident cells in the adult brain parenchyma. OPCs could generate oligodendrocytes to participate in myelination. Recent advances have renewed our knowledge of OPC biology by discovering novel markers of oligodendroglial cells, the myelin-independent roles of OPCs, and the regulatory mechanism of OPC development. In this review, we will explore the updated knowledge on OPC identity, their multifaceted roles in the CNS in health and diseases, as well as the regulatory mechanisms that are involved in their developmental stages, which hopefully would contribute to a further understanding of OPCs and attract attention in the field of OPC biology.

Keywords: Oligodendrocyte precursor cell, OPC-neuron synapse, Myelin-independent roles, Heterogeneity, Migration, Proliferation, Differentiation

Introduction

Oligodendrocyte precursor cells (OPCs, also known as NG2 glial cells) were first described over half a century ago, which were initially termed as either glioblasts, early oligodendrocytes, young oligodendrocytes, immature oligodendroblasts, or oligodendrocyte-type II astrocyte (O2A) progenitor cells, but were later identified and termed as oligodendrocyte progenitors or precursors [13]. OPCs are now considered the fourth major type of glial cells [4, 5], constituting 5–8% of glial cells in the adult central nervous system (CNS) [6].

OPCs have received increasing attention over the past years being recognized as a critical component in CNS health and diseases. OPCs can differentiate into oligodendrocytes (OLs) to myelinate axons during development and in adulthood [7] while generating Schwann cells during the repair of CNS demyelination and astrocytes in certain CNS regions [810]. OPC is also the only glial cell type in the CNS that forms synapses with neurons [1113], which impacts OPC developmental events such as migration, proliferation, and differentiation [12, 1418]. Recent advances also uncover the myelin-independent roles of OPCs, including the regulation of neuronal development and activity, astrocyte maturation, vascular formation, and function, as well as neuroinflammation [13, 1924]. These discoveries expand our knowledge of the multifaced role of OPCs, cultivating novel research areas in OPC biology. Additionally, with the up-and-coming new technologies, our understanding of the OPC developmental mechanism is also renewed, including the extrinsic and intrinsic regulation of OPC migration, proliferation, and differentiation.

In this review, we will cover the progress in the primary issues surrounding OPCs, including: (1) What are OPCs? (2) The function of OPCs. (3) OPC heterogeneity. (4) The regulatory mechanisms of OPC development. We hope to provide an outline for new researchers interested in this field.

What are OPCs?

The definition of OPCs is still in development since their discovery. The existence of OPCs was first suggested by ultrastructural electron microscopy studies between the 1960s and 1970s [2] and was further identified with the discovery of their markers, NG2, PDGFRα, and A2B5, in primary cultures and in vivo [2527]. Since then, a more holistic understanding of OPC identity, characteristics, and functions, has been reached through complex biochemical and genetic manipulation, as well as lineage tracing techniques. It is now considered that OPCs as neuroectoderm-origin multipotent progenitors that are NG2/PDGFRα positive, highly ramified, mainly generating OLs, but are also capable of giving rise to other CNS cell types including astrocytes.

Derivation of OPCs

During early embryonic development, oligodendroglial lineage cells are derived from radial glia, which are multipotent neural stem cells near the ependymal cell layer. They generate OPCs through asymmetric division [28], and OPC specification is regulated by several transcription factors, most notably Olig2 and Sox10, which together define the oligodendroglial lineage [29]. OPCs initially emerge in the CNS during middle gestation from Olig2+ neuroepithelial progenitors found in the ventricular zones of the spinal cord and brain [7]. During this specification progression, OPCs emerge in three originating waves from distinct domains in different regions of the forebrain, firstly from the medial ganglionic eminence at embryonic day (E)12.5, then the lateral ganglionic eminence at E15.5, and finally the dorsal cortex at birth [30, 31]. They then migrate extensively throughout the whole CNS along the brain vasculature as their guiding scaffolds [32]. Furthermore, they can also emerge indirectly through intermediate progenitors, such as glial restricted precursors (GRP), and potentially from other types of immature neural cells, including OPC pre-progenitors or pre-OPCs [3335]. These pathways illustrate the complex and multifaceted origins of OPCs within the CNS, highlighting their crucial role in the development and maintenance of neural networks. The oligodendrogenesis process is not limited to prenatal stages but continues postnatally and into adulthood, with new OPCs being generated from neural stem cells in the forebrain's subventricular zone (SVZ) [3639], or from Nestin+/CD13+ immature pericytes/mesenchymal cells in meninges and perivascular regions [40], upon demyelinating insults or ablation of OPCs.

Markers and Morphology

Various markers have been identified to label OPCs in different developmental states. PDGFRα and NG2 have been used to identify OPCs [27, 4143], but recent evidence has challenged their veracity as exclusive OPC indicators. For example, NG2 is also a marker of pericytes and is expressed by reactive macrophages/microglia following injury, leading to potential misinterpretation [4447]. Labeling with PDGFRα suffers from similar complications. PDGFRα, an essential surface receptor involved in OPC proliferation, and migration and as a regulator for timely differentiation [4850], is also expressed by vascular and leptomeningeal cells (VLMCs) [51]. Nonetheless, NG2 and PDGFRα remain the most reliable and widely used markers for OPCs, which could reveal their ramified morphology, a characteristic distinguishing OPCs from NG2-positive pericytes or PDGFRα-positive VLMCs, which display a flat morphology and are lack of processes.

In addition, transcription factors NKX2.1 and DLX2 can label early OPCs in the human telencephalon [52], while O4 labels the non-migratory, late OPCs [53]. The arrival of the single-cell era also provides novel markers for identifying OPCs as well as different stages of oligodendroglial lineage cells. In the developing human brain, single-cell RNA seq identified a transitional cell type lying between multi-potent neural progenitor cells and lineage-committed early OPCs, defined as “pre-OPC”, which possesses expression of both neural progenitor cell markers including GFAP, VIM, NES, HES1, NOTCH2, and EGFR, and oligodendroglial lineage markers OLIG1, OLIG2, and PDGFRα [54]. The same study also identified a potential novel OPC marker PCDH15, which regulates daughter cell separation after division [54]. In addition, differentiation-committed OPCs were also identified, with a unique expression pattern of Tcf7l2, Itpr2, Tmem2, Gpr17, and Pdgfa [51]. Among them, TCF7L2, ITPR2, and GPR17 are adopted as a marker for differentiation-committed OPCs or newly-formed OLs [51, 55]. In addition, activated OPCs in CNS injury highly express Rnf43, which is then downregulated during OL differentiation; RNF43 is thus considered a marker of reactive OPCs [56].

OPCs display variable morphologies, as reflected by the immunofluorescent staining of PDGFRα (Fig. 1). In early development (embryonic stage till birth), OPCs exhibit an elongated shape with bipolar processes during migration along brain vasculature [32]. Later (at around postnatal day 7 (P7)), they change to a branched morphology upon arrival at their destination, interacting with surrounding neural axons and other glial cells [5]. Live imaging using two-photon microscopy showed that adult OPCs possess elegantly ramified but slightly less branched processes (compared to OPCs at P7) in the mature CNS, where they constantly survey the environment within their territories to respond to local changes in tissue homeostasis [57]. Upon CNS injuries, OPCs either convert to a hypertrophic morphology with increased cell processes and upregulated NG2 expression level, or return to a bipolar migratory shape to facilitate their recruitment into lesion areas [22, 58, 59].

Fig. 1.

Fig. 1

Morphology changes of OPCs. A PDGFRα immunofluorescent staining and skeletonized image (right panel) visualize the altered OPC morphologies at different developmental stages. At postnatal day 0 (P0), most OPCs had a migratory morphology with some short and sparse protrusions (A1). At P7, the morphology of OPC processes was more complex, with more fine main protrusions and longer branches (A2). At P28, OPC protrusions and branches decreased (A3). Stars mark OPC soma. Immunostaining images, for illustration purposes, are courtesy of the laboratory of J.N. B Both migratory morphology OPCs with simple protrusions (B1) and reactivated OPCs with complex protrusions (B2) are visualized after lysolecithin-induced demyelinating injury at 3-day post lesioning (3 dpl). Stars mark OPC soma. Immunostaining images, for illustration purposes, are courtesy of the laboratory of J.N. C Schematics of OPC morphological changes.

The Function of OPCs

OPCs are a unique type of glial cell in the CNS, which could replenish the OL population for myelination or remyelination, transdifferentiation into astrocyte and Schwan cells, and could form synaptic connections with neurons. Recent advances also reveal the myelin-independent roles of OPCs. These characters further distinguish them from other glial cells, neurons, and neural progenitor cells [4, 6, 60].

Dynamic Architects in CNS Regeneration and Plasticity

OPCs are multipotent progenitors. While they mainly differentiate into OLs, they are also capable of generating astrocytes or transforming into a Schwann cell-like state in development and/or in CNS injury (Fig. 2A–D).

Fig. 2.

Fig. 2

Origin, function, and heterogeneity of OPCs. A Developmentally, OPCs differentiate into OLs and form myelin that enwrap axons. B and C Some OPCs maintain their undifferentiated state into adulthood, become adult OPCs, and populate the entire adult CNS. Adult OPCs continue to generate OLs and perform remyelination under physiological and pathological demyelination conditions. D Adult OPCs also have the potential to differentiate into astrocytes in diseases. E OPCs are the only glial cell types that can form synaptic connections with neurons, including the neuron-OPC synapse as well as the OPC-neuron synapse. F OPCs display regional, sex, and age heterogeneity in the CNS.

(Re)myelination

During development, OPCs proliferate and differentiate into OLs to perform developmental myelination [7]. However, around 20% of OPCs remain in an undifferentiated state throughout the developmental process and convert to adult OPCs in the mature CNS, where they act as multifaceted regulators of CNS health and disease [6163]. Many questions remained open about adult OPCs, including the difference between developmental OPCs and adult OPCs, and how adult OPCs maintain a quiescent, multipotent state. On the latter issue, it was proposed that contact with vasculature may participate in the maintenance of adult OPCs, as vascular endothelial cells are inhibitory for OPC differentiation [64]. Adult OPCs continue to generate OLs throughout life, albeit the rate decreases with aging [7]. While the innate OPC differentiation and myelination could take place independent of neuronal activity, accumulating evidence showed that neuronal activity could remodel myelin formation, a process called adaptive myelination, which involves changes in the thickness, length, and/or number of myelin sheaths on the axons [65, 66]. OPCs are actively involved in adaptive myelination via de novo OL generation [67]. This dynamic change of myelination in turn regulates signal transmission in neural circuits and implicates emotions and cognitive functions [68, 69].

OPCs also play a pivotal role in remyelination in demyelination injury. OPCs can sense myelin damage-related signaling and migrate into demyelinated lesions [70]. Genetic fate-mapping showed that upon demyelination by local injection of lysolecithins in the corpus callosum, parenchymal OPCs were rapidly recruited into demyelinating areas within 7 days and produced mature OLs within 14 days [71]. Interestingly, adult neural stem cells in the SVZ also contribute to remyelination in the adjacent demyelinated regions. Adult neural stem cells transit into OPCs and rapidly differentiate into OLs to aid remyelination [3739]. However, compared to the parenchymal OPCs, neural stem cell-derived OPCs respond to the demyelination injury more slowly and have less contribution to the de novo OL generation [71, 72].

Another controversial issue surrounding remyelination is whether adult OLs also participate in the remyelinating process [7375]. In an attempt to solve this, a study investigated the source of remyelinating OLs in human patients with MS using a 14C incorporation-based technique but showed that no new OLs were detected in remyelinating lesions [76], which is consistent with a previous non-human primate study [77]. While these studies suggest that residual OLs in the lesions may participate in remyelination, a recent report showed that compared to new OLs, the survived OLs have defective myelination regeneration, as they displayed limited myelination capacity [78] and could misguidedly myelinate neuronal cell body [79]. Meanwhile, several other studies failed to observe the contribution to remyelination from the pre-existing OLs in the galactocerebroside-antibody plus complement-induced demyelination model [80] or the lysolecithin-induced demyelination model [81]. Overall, these studies support the importance of OPCs in the remyelination process. However, one of the critical issues in human demyelination diseases is the failure of OPC migration into the demyelinated site [22, 82]. Therefore, stimulating OPC recruitment should benefit remyelination in demyelinated lesions.

Transformation and Conversion

It has long been proposed that OPCs have the potential to transdifferentiate into astrocytes in primary culture [26], hence OPC was initially named oligodendrocyte-type II astrocyte (O2A) progenitor cells [1]. However, observation in vivo using genetic fate mapping revealed that the transdifferentiation potential of OPCs is limited, and can only take place during early development or in CNS injury. Embryonic OPC transdifferentiation is more prominent in the ventral brain, while it is absent in the dorsal cortex [8]. This plasticity of OPC is regulated by transcription factor Olig2: knockout of Olig2 in NG2+ OPCs resulted in their transdifferentiation into ALDH1L1+ astrocytes [8]. Such Olig2-regulated OPC plasticity declines as mice reach adulthood [83]. In addition, engrafted human fetal OPCs, or a murine cerebrum OPC cell line, in the postnatal mouse brain could also differentiate into both OL and GFAP+ astrocytes [84, 85].

In CNS injury, OPCs can be regulated by signal molecules and transformed into astrocytes [1]. In a chronic ischemia model that yields severe myelin loss, BMP4-derived from pericytes promoted OPC transformation into astrocytes, which was proposed to hinder remyelination [86]. In the traumatic brain injury model, OPCs gave rise to astrocytes within 24 h post-injury but mainly differentiated into OLs to contribute to remyelination after 7 days [87]. The role of OPC-astrocyte conversion in CNS injuries is under debate. On the one hand, transdifferentiation of OPCs was proposed to hinder the process of remyelination, and targeting OPC transdifferentiation could be an effective strategy for the treatment of demyelination-related diseases and injuries [86]. On the other hand, OPC transdifferentiation might aid the formation of a glial barrier surrounding the injured niche that limits the spread of inflammation signal, which would facilitate post-injury tissue repairment [88, 89].

In spinal injury, OPC could also transdifferentiate into a Schwann cell-like state, expressing Schwann cell markers such as Periaxin (P0), and participating in myelin regeneration in the CNS [9, 90]. However, the identity of these Schwann cell-like states, as well as the regulatory mechanism, await further studies.

While it is recognized that OPC transdifferentiation into astrocyte and Schwann cells implicates disease progression, it remains unclear how OPC possesses the ability to transdifferentiate to astrocyte or Schwann cells. A combination of extracellular signals and intracellular regulation, e.g., epigenetic mechanism, might be at play to maintain OPC in a multipotent state, which would switch to the astrocyte- or Schwann cell fate upon pathological insult.

Neuron-OPC Synaptic Connection

OPC is the only resident glial cell type that forms direct synaptic connections with neurons in the CNS (Fig. 2E). Immuno-electron microscopic analysis revealed that silver-intensified gold reaction-labeled OPC processes approached neuronal dendritic spines and formed thinner postsynaptic membranes, while electrophysiological evidence confirmed a synaptic connection between glutamatergic neurons and OPCs [11]. The synaptic input from excitatory neurons to OPCs promotes myelination [14]. Meanwhile, OPCs also form inhibitory synapses with GABAergic interneurons [12]. GABAergic input to OPCs transiently suppresses the AMPA receptor-mediated glutamatergic current [12]. Additionally, functional inhibitory synapses between GABAergic interneurons and OPCs regulate various features of OPC development, including survival, proliferation, differentiation, myelination, and the length of CNS internodes [12, 1518]. Recently, a study utilizing a novel monosynaptic retrograde tracing technique further revealed that local OPCs receive inputs from multiple brain regions, which may promote myelination to support brain-wide circuits [14, 91]. The post-synaptic elements in OPCs gradually reduce when they differentiate into OLs [92]. It remains a puzzle why both glutamatergic and GABAergic input promote OPC differentiation and myelination, and if other neuronal cell types, e.g., cholinergic or dopaminergic neurons, could form a synapse with OPC, and thus regulate OPC differentiation.

More importantly, why OPC is necessarily the only glial cell type that forms synapses with neurons, instead of just reacting to spillover neurotransmitters like other glial cells, such as astrocytes [93, 94]? From a parsimonious viewpoint, neuron-OPC synapse should bear physiological significance. It may be required for the precisive myelin formation, which is hinted by a recent study in the zebrafish model, showing that GABAergic post-synaptic hotspots in OPCs predict a subset of future myelin sheath formation sites [92].

Apart from the neuron-OPC synapse, recent research also suggested that OPCs can form functional GABAergic synapses onto neurons in the hippocampus of adult mice, and a single OPC is capable of forming approximately 146 presynaptic structures per cell for signal transduction to neurons in the hippocampal CA1 region [13], and receive inputs from 141 different axons in white matter [95]. Bidirectional communication between OPCs and interneurons has since been also discovered, where GABAergic interneurons regulate OPC myelination, and OPCs can also shape the inhibitory neuronal network through GABA receptors and the cytokine tumor necrosis factor-like weak inducer of apoptosis (TWEAK) signaling pathways [13, 19].

Myelin-Independent Functions

The last few years have witnessed a surge of research uncovering the myelin-independent roles of OPCs in health and diseases, which have been well-summarized in several reviews recently [9699].

OPCs have been shown to modulate neuronal migration, axon arborization, and neural activity, and participate in synapse engulfment and synaptogenesis. Developmental OPC ventral-dorsal perivascular migration acts to steer the migration of their predecessor, the ventral-originated interneurons, away from the blood vessel to facilitate their dispersion to the dorsal brain [100]. OPC also regulates the arborization of axons and fine-tuning the neural circuit formation [101]. The aforementioned OPC-neuron synapse promotes the tonic GABA release to promote inhibitory synapse transmission [13]. The neuron-OPC synapse, on the other hand, may have a reciprocal effect on neurons, as conditional knockout of GABA receptor subunit GABABR in OPC using the NG2-CreERT line resulted in reduced inhibitory neuron apoptosis, leading to an increased inhibitory neuron density in the prefrontal cortex [19]. Electron microscopic evidence shows that OPC fine processes are frequently in touch with axons and contain numerous phagosomes, which suggests an axon-trimming role of OPCs [102]. In vivo, phagocytosis assay confirms that OPC could engulf synapses to participate in circuit remodeling [103]. In a mouse model of schizophrenia, pathological OPCs suppressed hippocampal neuron synaptogenesis via secretion of WIF1 [104].

OPCs also regulate other brain cell types and implicate neuroinflammation in disease conditions. During development, OPC-derived Wnt ligands Wnt7a/7b promote angiogenesis [20] as well as the maturation of the astrocyte network [21]. In demyelinating diseases, OPCs could directly interact with the vasculature and interrupt the blood-brain barrier function, aggravating the neuroinflammatory condition [22]. In addition, OPC is not only a “receiver” of inflammatory signals, but also an “instigator”: while OPCs respond to inflammatory cytokines or chemokines, which may disrupt the remyelination process, reactive OPCs could also release immune factors such as CCL2 and IL1β, and present antigens, which may participate in the recruitment of peripheral immune cells and promote the inflammatory response [23, 24].

The myelin-independent role of OPC and the underlying mechanism remain largely unexplored and are becoming an exciting field for future studies.

OPC Heterogeneity

OPCs are a heterogeneous glial cell population with various functions both in development and in the adult CNS [105108] (Fig. 2F).

Regional Heterogeneity

OPCs exhibit heterogeneity across different brain regions. One important evidence comes from the radiation-mediated OPC-depletion experiments, where OPCs in different brain regions show diverse radiation resistance and post-radiation proliferative capability. In the cortex, corpus callosum, and hippocampus, OPCs were almost completely depleted 7 days post-radiation, followed by a progressive repopulation from the non-irradiated area [109]. In contrast, OPC loss occurred slower in the central regions of the brain (midbrain and hypothalamus), with 26% of OPCs still present 4 weeks after radiation [109]. Transplanted and endogenous OPCs quickly established in the depleted cortex, but not in the partially depleted central regions, which suggests a degree of regional heterogeneity [109]. A similar heterogeneous response to stress was also observed in a DNA damage model by ablation of citron-kinase or cisplatin treatment [110]. OPCs in the dorsal brain are more susceptible to DNA damage and rapidly undergo apoptosis, while the ventral ones only display senescence phenotype [110]. In demyelination models, dorsal OPCs outperformed their ventral counterparts in proliferation, recruitment, and differentiation [111, 112]. A recent lineage tracing study in mice made a surprising discovery that the dorsal-originated OPC population continuously expanded throughout life, while the ventral-originated one gradually diminished; at 6 months of age, dorsal-originated OPCs can account for almost all OPCs in the brain [61].

There is still controversy over the origin of such regional OPC heterogeneity. Some argue that it stems from the difference in the developmental origin of dorsal and ventral OPCs [61, 111, 112]. On the contrary, it is also proposed that the local environmental cues are the main contributors [110], which was well supported by the aforementioned OPC transplantation experiment [109]. In addition, while the behavioral differences between OPC in dorsal and ventral regions are well documented, their molecular bases await further studies.

OPC heterogeneity was also observed between grey matter and white matter, displaying different morphology and responsiveness to PDGF [113]. A study in the zebrafish model further illustrates how neuronal-rich area and axon-rich area shape OPC heterogeneity, characterizing the difference in transcriptome, morphology, and differentiation kinetics [108]. Morphology-wise, OPCs with somata residing in the neuron-rich area form a more elaborate process network than those within the axon-rich area [108]. Moreover, live-imaging showed that OPCs within the axon-rich area had higher process dynamics, as they completely remodeled their entire process network in 1 h, whereas their neuron-rich area counterparts only remodeled the process tips [108]. Differentiation-wise, OPCs in the axon-rich area tend to differentiate rapidly into myelinating OLs, while OPCs in the neuron-rich area proliferate in response to neural activity, and migrate to plenish the axon-rich area OPC pool [108]. These observations in zebrafish reveal how the microenvironments in the neuron-rich grey matter and axon-rich white matter may shape OPC heterogeneity.

Age Heterogeneity

Gene tracing studies indicate that the age-related decrease in the number of OPCs may result in reduced myelin renewal and memory impairment in mouse models [114]. Single-cell electrophysiological recordings showed that OPC heterogeneity becomes more prominent with age, even though they start as a relatively homogeneous population. In immunostaining studies, a subset of OPCs lacking OLIG2, ranging from 2 to 26%, has been identified in various brain regions including the cortex, corpus callosum, CA1, and dentate gyrus. These OLIG2-negative OPCs are enriched in the young brain and decrease with age and are rarely detected in the adult brain [115]. Single-cell transcriptome analysis also revealed that OPC molecular signatures change with age, which is reflected by that embryonic OPCs possess a higher expression of migrating signatures (expressing genes such as DCC and EPHB2) compared to OPCs from postnatal or older mice. OPCs within 2 weeks of life had stronger proliferative signatures (for example, PDGFRα and PTCH1) but reduced migratory molecular signatures. Molecular signatures of both differentiation and proliferation begin to decline from 2 weeks after birth to the following months and decline further in aged mice [116]. In another single-cell transcriptome dataset, four OPC cell clusters were identified that distinguish young and old mouse OPCs. Some subsets of OPC clusters are closely related to cell division and differentiation, but their numbers decrease in the aging brain, in-depth analysis of single-cell RNA sequencing data from all brain cells from young and old mice revealed upregulation of several aging-related genes in a subset of OPCs in the aging brain [117].

Sexual Heterogeneity

OPCs exhibit differences between genders in terms of proliferative and differentiative potential, metabolism, regional heterogeneity, and response to pathological insults. In primary culture, the proliferation and migration ability of OPCs from female animals is higher, whereas the differentiation ability of OPCs from male animals is stronger [118]. Similar differences in proliferation and differentiation capability can be seen in OPCs from different genders in vivo [118, 119], suggesting that the ability of OPCs to proliferate and differentiate is closely related to gender. OPCs exhibit gender-related differences in energy metabolism: the intracellular ATP level is significantly higher in female OPCs than in male OPCs [118]. A recent single-cell RNA sequencing study has shown that male OPCs are more abundant in the interventricular septum than females, indicating gender-specific regional differences in the adult brain [120].

Gender differences in OPCs may also play an important role in the susceptibility and response to neurological and psychiatric diseases. In Alzheimer's Disease (AD), OPCs and OLs show gender-specific differences in the transcriptomic response to AD pathology. Single-cell RNA sequencing from post-mortem human brain tissue shows that OLs from males upregulate transcriptional genes in pathological reactions, while OLs from female patients do not [121]. Additionally, OPCs from females tend to downregulate transcriptional genes in pathological reactions compared to their male counterparts [121]. It is suggested that this may be partially due to differences in OPC proliferation and differentiation caused by sex hormones or biological gender [1]. Similarly, OPCs have gender-specific differences in cell death in response to ischemic stress. Male OPCs are more susceptible to in vitro oxygen-glucose deprivation. After 7 h of oxygen-glucose deprivation (OGD), the cytotoxicity and activity of OPCs were evaluated in male OPCs, and measurements of the cytotoxicity ratio (cytotoxicity/activity) showed that the cytotoxicity of male OPCs is significantly higher than that of female OPCs after OGD [118]. As gender-dependent OPC responses to neurological conditions correlate with the gender-difference in disease-susceptibility, the related mechanistic investigation would yield valuable insight into disease pathogenesis.

The Regulatory Mechanisms of OPC Development

OPC development involves several critical steps, including migration, proliferation, differentiation, and finally myelination. These steps tightly connect, and are precisely regulated (Table 1).

Table 1.

Molecular regulations for OPC development

Development Role Type Factor Evidence
Migration Attraction Extracellular factor/signaling PDGF [129132]; FGF [132]; SEMA3F [137139]; HMGB1 [140]; SHH [123] In vitro
Repulsion Extracellular factor/signaling BMP [122] In vitro
SEMA3A [137139] In vitro; In vivo
CXCL1 [133]; Netrin1 [134, 135] In vitro; Ex vivo; In vivo
Proliferation Promotion Extracellular factor/signaling PCDH15 [54] Ex vivo
FGF2 [146, 149]; IGF1 [147]; GGF [150] In vitro
BDNF [148]; NT3 [148, 149]; PDGF [48, 144, 145, 149] In vitro; In vivo
Channel proteins Cx29, Cx47 [143] In vitro
Transcription factors ID2, ID4 [172, 196198]; SOX2, SOX5, SOX6 [199203]; SRF [153] In vitro
Epigenetic regulation SIRT1 [214]; SIRT2 [215] In vitro; In vivo
Suppression Extracellular factor/signaling TGF-β [151]; BMP [152] In vitro
Transcription factors ASCL1 [193] In vivo
Differentiation Promotion Extracellular factor/signaling IGF-1 [164, 165]; CNTF [166]; FGF [167] In vitro
Transcription factors OLIG2 [8, 175, 176, 178, 179]; OLIG1 [182]; SOX10 [183, 184]; MYRF [187]; ZFP24 [190, 191]; PPARγ [206]; RXR [207] In vitro; In vivo
TCF7L2 [156, 192]; ASCL1 [193]; FOXG1 [204, 205] In vivo
Epigenetic regulation SETDB1 [179]; CHD7, CHD8 [208210]; DNMT1 [219]; TET1 [222]; DICER1 [223, 224] In vivo
PRMT5 [211213]; SIRT2 [215]; miR-219, miR-338, miR-138 [223225]; miR-146a [226]; miR-21-5p [227] In vitro; In vivo
Suppression Extracellular factor/signaling NOTCH [168, 169]; Wnt/β-catenin [156159]; BMP [172174] In vivo
Transcription factors HES1, HES5 [194, 195]; ID2, ID4 [172, 196198]; SOX2, SOX5, SOX6 [199203] In vitro
Epigenetic regulation SIRT1 [214] In vitro; In vivo

Regulation of OPC Migration

OPCs migrate throughout the CNS and are guided by a variety of molecules, such as growth factors, morphogens [122, 123], neurotransmitters or neuromodulators [124126], as well as extracellular matrix [127, 128]. Chemo-attractive and chemo-repulsive factors co-ordinate to regulate OPC migration and intricately determine their positions. For instance, PDGF [129132], FGF [132], and SHH [123] represent a chemo-attractive force, while BMP [122], CXCL1 [133], and Netrin1 [134, 135] repel OPCs to direct ventral-dorsal migration, as well as to stop OPC migration when reaching their destination. Semaphorins (SEMA) are a family of membrane proteins that primarily regulate cell chemotaxis [136]. In development and demyelination diseases, it was shown that SEMA3A suppresses OPC migration while SEMA3F promotes OPC migration [137139]. In demyelinating injury, High mobility group box-1 (HMGB1) acts as an autocrine chemo-attractant for OPC migration [140].

Recent advances highlighted the critical contribution of cell-cell interaction in the delicate control of OPC migration. In the brain or spinal cord, migrating OPCs use blood vessels as physical scaffolds to rapidly disperse throughout the neural tissue or toward demyelinated lesions [22, 32]. Later in development, the peri-vascular OPC migration is ended by the formation of an astrocyte endfeet that suppresses the interaction between OPC and vasculature [64]. However, further studies are required to address the distance and speed of OPC migration, as well as any differences between migration during development and regeneration.

Regulation of OPC Proliferation

Various factors influence the proliferation of OPCs during their development. OPC density is one of the critical regulators of OPC proliferation. Early evidence showed that in primary culture, high seeding density suppressed OPC proliferation [141]. OPCs in vitro display interesting properties that after cell division, the daughter cells rapidly repulse each other and migrate to the opposite direction, distinct from the LIFR+ or EGFR+ neural progenitors that tend to stay in contact with each other after division [54]. This self-repulsion of OPC is mediated by a membrane-located protocadherin protein PCDH15; failure of self-repulsion in PCDH15 KO OPC is accompanied by a dampened OPC proliferation [54]. These data suggest OPC density exerts a contact-inhibitory effect on OPC proliferation. Such balance between cell growth and self-repulsion persists till adulthood to maintain the OPC population [57]. However, our knowledge of the underlying mechanism remained rudimentary.

OPC proliferation is also sensitive to energy supply. OPC-expressed connexin proteins Cx29 and Cx47 form hemichannel and absorb energy substances such as glucose to maintain cell proliferation [142, 143]. During hypoglycemia, brain levels of high-energy phosphate molecules decrease, cell membranes release free fatty acids, intracellular calcium levels increase and membranes depolarize. We and other researchers have observed that the proliferation and migration of OPCs are inhibited at low glucose levels, suggesting that OPCs utilize glucose as an important energy source for growth and motility [142, 143].

OPC proliferation is dynamically regulated by growth factors or morphogens. Growth factors that promote OPC proliferation include PDGF [144, 145], FGF2 [146], IGF1 [147], brain-derived neurotrophic factor (BDNF) [148], neurotrophic factor 3 (NT3) [149], and glial cell-derived growth factor (GGF) [150]. Factors that inhibit OPC proliferation include transforming growth factor-beta (TGF-β) [151] and BMPs [152]. However, most of these conclusions are derived from in vitro data. The in vivo contribution of different growth factors in OPC proliferation during development and diseases required further investigation.

Transcription factors and epigenetic mechanisms also play a regulatory role in the proliferation of OPCs. Transcription factors such as SRF (serum response factor) [153] promote the proliferation of OPCs. The epigenetic mechanisms related to the proliferation of OPCs mainly include DNA methylation, histone modification, and non-coding RNA. These mechanisms can influence gene expression and thus regulate the proliferation of OPCs [154].

Regulation of OPC Differentiation

Extracellular Cues and Signaling Pathways

Extracellular factors play important roles in OPC differentiation via activation of intracellular signaling pathways. These include the Wnt/β-catenin signaling pathway, in which Wnt ligands act on the transmembrane Lipoprotein receptor-related protein (LRP)-Frizzled receptor complex, leading to the stabilization and nuclear translocation of β-catenin, which then binds with transcription factor TCF to activate gene transcription [155]. The Wnt/β-catenin signaling pathway plays a complex role in OPC differentiation. Hyper-activated Wnt/β-catenin signaling would perturb OPC differentiation [156159]. However, the Wnt/β-catenin signaling pathway is required for the expression of TCF7L2 [160, 161], a transcription factor and an effector of Wnt/β-catenin signaling, which propel OPC differentiation and myelination. One key mechanism involves the induction of Axin2 by Wnt/β-catenin-TCF7L2 signaling, which suppresses Wnt signaling as a negative feedback loop, allowing subsequent OPC differentiation [162, 163].

Several extracellular factors have been shown to promote OPC differentiation. Insulin-like growth factor-1 (IGF-1) could suppress BMP signaling to instruct oligodendroglial lineage specification from neural progenitor [164], while in OPCs, IGF-1 and extracellular matrix protein Laminin cooperate to activate the MAPK pathway in the promotion of OPC differentiation [165]. In addition, ciliary neurotrophic factor (CNTF) is a neurotrophic factor produced by neurons and glial cells, which can promote OPC differentiation into mature OLs through the gp130-JAK signaling pathway [166]. FGF was proposed to be a “late stage” signal, acting on FGF receptor 2 of OLs to regulate myelin thickness through activation of the MAPK-mTOR pathway [167].

On the other hand, the Notch signaling pathway suppresses OPC differentiation and maintains the OPC state. At the early stage of OPC development, JAGG1(or Jag-1) is highly expressed on axonal membranes and can activate the NOTCH receptor on OPCs, leading to the cleavage of the NOTCH receptor and the nuclear-translocation of its intracellular domain, which promotes the expression of transcription factor HES5, and inhibit OPC differentiation and myelin formation [168, 169].

BMP signaling is another important pathway in regulating OPC differentiation. BMPs signal through serine/threonine kinase receptors on the membrane, leading to the phosphorylation and nuclear-translocation of SMAD proteins and the subsequent transcription regulation [170]. Activation of BMP signaling in OPCs by BMP2/4 causes transdifferentiation to astrocytes via an upregulation of transcription factor ID2/4 and a downregulation of OLIG1/2 [171173].

Transcription Factor-Mediated Regulation

The transcription factor regulatory network that controls OPC differentiation has been intensively studied, revealing several key regulatory factors, with OLIG2, SOX10, Nkx2.2, ZBP24, and MYRF being the major determinants [174]. OLIG2, a basic helix-loop-helix transcription factor, is a master regulator of oligodendrocyte lineage cells. OLIG2 ablation in OPCs using the NG2-CreERT line during development causes loss of OPC marker expression and the transdifferentiation to astrocyte, as well as the subsequent hypomyelination [8, 175], while overexpression of OLIG2 enhanced OPC differentiation [176]. Mechanistically, OLIG2 interacts with Nkx2.2 [177] and engages with SWI/SNF chromatin remodeler SMARCA4 (also known as BRG1) to promote gene expression that involves the specification of oligodendroglial lineage as well as OPC differentiation [178]. A recent report also reveals that OLIG2 could also recruit SETDB1 to achieve gene silencing, which is also essential for OPC differentiation and myelinogenesis [179]. However, Plp-CreER-mediated conditional knockout of OLIG2 in OLs facilitates myelination [180], possibly via the compensating upregulation of OLIG1 expression. OLIG1 is a transcription factors that show structural similarity with OLIG2; OLIG1 and OLIG2 display coordinated expression in OPC during development, and their functions are partially redundant [181]. Interestingly, as OPCs differentiate, phosphorylation of OLIG1 would cause its translocation from the nucleus to the cytoplasm, which is crucial for OL morphological maturation and myelination [182].

OLIG2 directly promotes the expression of SRY-box 10 (SOX10), a transcription factor crucial for OL maturation [183, 184], which in turn forms a positive feedback loop to maintain OLIG2 expression [185]. SOX10 is required for expression of myelination-related genes, including transcription factors MYRF and ZFP24; a complex positive feedback network among OLIG2, SOX10, MYRF, and ZFP24 is crucial for normal OPC differentiation [178, 186191].

Transcription factor TCF7L2 is a recently identified newly-formed OL marker. It is an effector of Wnt/β-catenin signaling and interacts with transcription co-repressor ZBTB33 to block β-catenin signaling, a negative feedback regulation that is required for differentiation and myelination [156]. TCF7L2 also cooperates with SOX10 to promote the expression of MYRF and ZFP24 [192].

Apart from this OPC-specific regulation, transcription factors that have been at play since the neural progenitor stages also implicate OPC differentiation. This includes ASCL1 which balances OPC proliferation and differentiation: Pdgfrα-CreERT2-mediated conditional knockout of ASCL1 promoted OPC proliferation and impeded differentiation [193]. Transcription factors HES1 and HES5 are effectors of the Notch signaling pathway, which negatively regulate ASCL1 expression and suppress OPC differentiation [194, 195]. ID2 and ID4 transcription factors act downstream of BMP and GPR17 signaling to promote OPC proliferation and suppress OPC differentiation, possibly via the sequestration of OLIG1/2 [172, 196198]. In addition, SOX2/5/6 inhibits OPC differentiation and promotes proliferation [199203]. Recent research showed that forkhead box transcription factor FOXG1 in OPC promotes cell cycle exit and differentiation in development and demyelinating injury [204, 205].

OPC differentiation is also regulated by nuclear receptors that are ligand-activated transcription factors, such as PPARγ and RXR, which are required for differentiation and myelination [206, 207].

Epigenetic Regulation

Transcription factors team up with epigenetic modifications to regulate OPC differentiation, which include chromatin remodeling, histone modifications, DNA methylation, and non-coding RNA. Transcription factor Olig2 recruits chromatin modifiers such as the BRG1 complex to alter the chromatin-accessibility landscape, which promotes or silences target gene transcription [178]. Olig2 also recruits an H3K9 tri-methyltransferase SETDB1 to achieve gene silencing and promote OPC differentiation [179]. Several other chromatin remodelers such as CHD7 and CHD8 have been shown to associate with transcription factors such as SOX10 to promote OPC differentiation [208210].

Experiments were also performed to examine the role of histone modification enzymes in the regulation of OPC differentiation. PRMT5, which is a methyltransferase of histone 4 arginine 3 (H4R3) and histone 3 arginine 8 (H3R8) and achieves gene silencing, was proposed to promote OPC differentiation and myelination via suppression of ID2/4 expression in primary culture [211]. PRMT5 was further confirmed to be necessary for OPC survival and differentiation in vivo in Olig1-Cre or Olig2-Cre-mediated conditional knockout models [212, 213]. It was further shown that reduced histone arginine methylation by Prmt5 knockout was followed by increased histone lysine acetylation, and that defective myelination in Prmt5 knockout OPC could be rescued by histone acetyltransferase (HAT) inhibitor [212]. Histone acetylation is established by HATs, while acetyl removal is maintained by histone deacetylases (HDACs). SIRT1, the nicotinamide adenine dinucleotide (NAD)-dependent class III HDAC, promotes OPC proliferation while suppressing differentiation as suggested by the knockdown experiment [214]. In contrast, SIRT2 could promote OPC differentiation [215]. It is interesting that while both are histone deacetylases, SIRT1 and SIRT2 have opposite roles in OPC differentiation, which might be due to their alternative interactors [215]. In the adult brain, deactivation of SIRT1 in neural stem cells promotes the expansion of oligodendroglial lineage cells but does not alter OPC differentiation or myelination [216]. Interestingly, HDAC can also regulate OL development in a histone-independent manner, such as by the deacetylation of transcription factors [217, 218].

DNA methylation usually correlates with gene silencing and is dynamically regulated during OPC differentiation. Increased DNA methylation level was found in OLs compared to OPCs, accompanied by a decreased expression of DNA methyltransferase DNMT1; however, knockout of Dnmt1 in oligodendroglial cells by Olig1-Cre line causes a defective OPC differentiation and the subsequent hypomyelination [219]. It is proposed that the non-selected DNA hypomethylation by ablation of DNMT1 causes defective alternative splicing and the subsequent ER stress, thus leading to impeded OPC differentiation [219]. Thus, the role and mechanism of dynamic DNA methylation in OPC differentiation remain to be studied. It was also shown that differentiation inhibiting factors ID2/4 are one of the subjects under DNA methylation control, which reduces their expression to facilitate OPC differentiation [220]. Age-related myelin loss might be also regulated by DNA methylation, as the level of genomic DNA methylation and the activity of methyltransferases gradually decrease in spinal cord OPCs during aging [221]. DNA hydroxymethylation, another form of DNA modification, is an intermediate product generated during demethylation catalyzed by the ten-eleven translocation (TET) proteins. Interestingly, Tet1 ablation in oligodendroglial lineage cells also causes defective proliferation and differentiation, leading to hypomyelination in adolescent mice, which was then normalized in adulthood [222]. This, again, highlights the role of DNA methylation dynamics in OPC differentiation. The investigation of the interaction between transcription factors, chromatin remodelers, and DNA methyltransferases might yield beneficial insight into how DNA methylation dynamics are involved in the regulation of OPC differentiation.

miRNA is another level of epigenetic regulation of OPC differentiation. miRNA is a type of small noncoding RNA molecule with an average length of 21–25 nucleotides that can complementarily bind to the target mRNAs, thereby silencing their expression. The necessity of miRNA-mediated regulation on OPC differentiation was first examined by two pioneer studies focusing on the miRNA processing enzyme DICER1. Dicer1 conditional knockout in oligodendroglial cells by Olig1-Cre, Olig2-Cre, or Cnp-Cre lines causes severe OPC differentiation defect and hypomyelination while promoting OPC proliferation [223, 224]. Several DISC1-processed miRNAs, which were upregulated during OL differentiation and reduced by Dicer1 knockout, were then confirmed to promote OPC differentiation by gain-of-function assays, including miR-219 [223, 224], miR-338 [224], and miR138 [224]. The target transcripts of these miRNAs were further identified. miR-219 targets genes essential for maintaining OPC proliferation (such as SOX6, FoxJ3, ZFP238, and PDGFRα), its increased expression thus stimulates OPC exit from the proliferative cycle and entry into differentiation [223225]. miR-338 share several common targets with miR-219 including SOX6, and also target gene transcripts such as Mytl1 [223]. A follow-up study showed that miR-219 cooperates with miR-338 to promote myelination, as miR-338 knockout aggravates the hypomyelination of Olig1-Cre-mediated miR-219 conditional knockout mice, and further presents the therapeutic potential of miR-219 in models of multiple sclerosis [225]. Additionally, miR-146a and miR-21-5p were also shown to promote OPC differentiation in vitro [226, 227]. While these data substantiate the critical role of DICER1 and miRNAs in OPC differentiation and myelination, it remained a puzzle how the production of these miRNAs was dynamically regulated in OPCs.

Conclusion

OPCs play critical roles in CNS function and homeostasis. This review covers how OPCs participate in myelination and remyelination processes, transdifferentiate into astrocytes and Schwann cells, as well as the recently identified myelin-independent roles. We also discuss the heterogeneity of OPCs across different regions, ages, and sex. Finally, we review the molecular mechanism regulating OPC development. With the development of biochemical and genetic manipulation techniques, the overall understanding of OPCs has been improved. Nonetheless, our understanding of OPCs is still very rudimentary, with numerous open questions.

One key aspect is the physiological role of OPCs, including: How does adult OPC maintain multipotent and quiescent? Are the neuron-OPC synapses associated with precisive myelin formation? How are the OPC differentiation and myelination required for learning? What are the biological roles of OPC in CNS homeostatic support, beyond myelination?

The second aspect revolves around the pathological role of OPCs. What is the spectrum of OPC pathophysiological changes and how these are modified in different diseases or disease stages? What is the role of OPC transdifferentiation in demyelinated injuries? What prevents OPC differentiation and remyelination in demyelinated lesions and how can we overcome the obstacles? How do these pathological changes in OPC integrate into the nervous tissue in response to various pathological insults?

Only by answering all these questions will we be able to achieve a comprehensive understanding of OPC biology and pathophysiology, and hopefully will contribute to the development of OPC-oriented therapeutic strategies.

Acknowledgments

This review was supported by grants from the National Natural Science Foundation of China (32271034, 32070964, 81971309, 32170980, and 32300791); National Key Research and Development Program of China (2021ZD0201703); Chongqing Natural Science Fund for Distinguished Young Scholars (CSTB2023NSCQ-JQX0030); Guangdong Basic and Applied Basic Research Foundation (2022B1515020012, 2021A1515110268, 2023A1515010651), Shenzhen Medical Research Fund (A2303014), Shenzhen Fundamental Research Program (JCYJ20210324123212035, RCYX20200714114644167, ZDSYS20220606100801003, RCBS20210706092411028, and JCYJ20210324121214039), and Shenzhen Key Laboratory of Chinese Medicine Active Substance Screening and Translational Research (ZDSYS20220606100801003).

Conflict of interest

The authors declare that they have no conflict of interest with the contents of this article.

Footnotes

Zexuan Ma, Wei Zhang, and Chenmeng Wang have contributed equally to this review.

Contributor Information

Chenju Yi, Email: yichj@mail.sysu.edu.cn.

Jianqin Niu, Email: jianqinniu@tmmu.edu.cn.

References

  • 1.Akay LA, Effenberger AH, Tsai LH. Cell of all trades: Oligodendrocyte precursor cells in synaptic, vascular, and immune function. Genes Dev 2021, 35: 180–198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Privat A. Postnatal gliogenesis in the mammalian brain. Int Rev Cytol 1975, 40: 281–323. [DOI] [PubMed] [Google Scholar]
  • 3.Hirano M, Goldman JE. Gliogenesis in rat spinal cord: Evidence for origin of astrocytes and oligodendrocytes from radial precursors. J Neurosci Res 1988, 21: 155–167. [DOI] [PubMed] [Google Scholar]
  • 4.Peters A. A fourth type of neuroglial cell in the adult central nervous system. J Neurocytol 2004, 33: 345–357. [DOI] [PubMed] [Google Scholar]
  • 5.Nishiyama A, Komitova M, Suzuki R, Zhu X. Polydendrocytes (NG2 cells): Multifunctional cells with lineage plasticity. Nat Rev Neurosci 2009, 10: 9–22. [DOI] [PubMed] [Google Scholar]
  • 6.Levine JM, Reynolds R, Fawcett JW. The oligodendrocyte precursor cell in health and disease. Trends Neurosci 2001, 24: 39–47. [DOI] [PubMed] [Google Scholar]
  • 7.Bergles DE, Richardson WD. Oligodendrocyte development and plasticity. Cold Spring Harb Perspect Biol 2015, 8: a020453. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Zhu X, Zuo H, Maher BJ, Serwanski DR, LoTurco JJ, Lu QR. Olig2-dependent developmental fate switch of NG2 cells. Development 2012, 139: 2299–2307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Assinck P, Duncan GJ, Plemel JR, Lee MJ, Stratton JA, Manesh SB, et al. Myelinogenic Plasticity of Oligodendrocyte Precursor Cells following Spinal Cord Contusion Injury. J Neurosci 2017, 37: 8635–8654. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Bartus K, Burnside ER, Galino J, James ND, Bennett DLH, Bradbury EJ. ErbB receptor signaling directly controls oligodendrocyte progenitor cell transformation and spontaneous remyelination after spinal cord injury. Glia 2019, 67: 1036–1046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Bergles DE, Roberts JD, Somogyi P, Jahr CE. Glutamatergic synapses on oligodendrocyte precursor cells in the hippocampus. Nature 2000, 405: 187–191. [DOI] [PubMed] [Google Scholar]
  • 12.Lin SC, Bergles DE. Synaptic signaling between GABAergic interneurons and oligodendrocyte precursor cells in the hippocampus. Nat Neurosci 2004, 7: 24–32. [DOI] [PubMed] [Google Scholar]
  • 13.Zhang X, Liu Y, Hong X, Li X, Meshul CK, Moore C, et al. NG2 glia-derived GABA release tunes inhibitory synapses and contributes to stress-induced anxiety. Nat Commun 2021, 12: 5740. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Ziskin JL, Nishiyama A, Rubio M, Fukaya M, Bergles DE. Vesicular release of glutamate from unmyelinated axons in white matter. Nat Neurosci 2007, 10: 321–330. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Orduz D, Maldonado PP, Balia M, Vélez-Fort M, de Sars V, Yanagawa Y, et al. Interneurons and oligodendrocyte progenitors form a structured synaptic network in the developing neocortex. eLife 2015, 4: e06953. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Hamilton NB, Clarke LE, Arancibia-Carcamo IL, Kougioumtzidou E, Matthey M, Káradóttir R, et al. Endogenous GABA controls oligodendrocyte lineage cell number, myelination, and CNS internode length. Glia 2017, 65: 309–321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Zonouzi M, Scafidi J, Li P, McEllin B, Edwards J, Dupree JL, et al. GABAergic regulation of cerebellar NG2 cell development is altered in perinatal white matter injury. Nat Neurosci 2015, 18: 674–682. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Moura DMS, Brennan EJ, Brock R, Cocas LA. Neuron to oligodendrocyte precursor cell synapses: Protagonists in oligodendrocyte development and myelination, and targets for therapeutics. Front Neurosci 2021, 15: 779125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Fang LP, Zhao N, Caudal LC, Chang HF, Zhao R, Lin CH, et al. Impaired bidirectional communication between interneurons and oligodendrocyte precursor cells affects social cognitive behavior. Nat Commun 2022, 13: 1394. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Yuen TJ, Silbereis JC, Griveau A, Chang SM, Daneman R, Fancy SPJ, et al. Oligodendrocyte-encoded HIF function couples postnatal myelination and white matter angiogenesis. Cell 2014, 158: 383–396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Wang Y, Su Y, Yu G, Wang X, Chen X, Yu B, et al. Reduced oligodendrocyte precursor cell impairs astrocytic development in early life stress. Adv Sci 2021, 8: e2101181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Niu J, Tsai HH, Hoi KK, Huang N, Yu G, Kim K, et al. Aberrant oligodendroglial–vascular interactions disrupt the blood–brain barrier, triggering CNS inflammation. Nat Neurosci 2019, 22: 709–718. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Moyon S, Dubessy AL, Aigrot MS, Trotter M, Huang JK, Dauphinot L, et al. Demyelination causes adult CNS progenitors to revert to an immature state and express immune cues that support their migration. J Neurosci 2015, 35: 4–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Kirby L, Jin J, Cardona JG, Smith MD, Martin KA, Wang J, et al. Oligodendrocyte precursor cells present antigen and are cytotoxic targets in inflammatory demyelination. Nat Commun 2019, 10: 3887. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Stallcup WB, Beasley L. Bipotential glial precursor cells of the optic nerve express the NG2 proteoglycan. J Neurosci 1987, 7: 2737–2744. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Raff MC, Miller RH, Noble M. A glial progenitor cell that develops in vitro into an astrocyte or an oligodendrocyte depending on culture medium. Nature 1983, 303: 390–396. [DOI] [PubMed] [Google Scholar]
  • 27.Nishiyama A, Lin XH, Giese N, Heldin CH, Stallcup WB. Co-localization of NG2 proteoglycan and PDGF alpha-receptor on O2A progenitor cells in the developing rat brain. J Neurosci Res 1996, 43: 299–314. [DOI] [PubMed] [Google Scholar]
  • 28.Qian X, Shen Q, Goderie SK, He W, Capela A, Davis AA, et al. Timing of CNS cell generation: A programmed sequence of neuron and glial cell production from isolated murine cortical stem cells. Neuron 2000, 28: 69–80. [DOI] [PubMed] [Google Scholar]
  • 29.Sock E, Wegner M. Using the lineage determinants Olig2 and Sox10 to explore transcriptional regulation of oligodendrocyte development. Dev Neurobiol 2021, 81: 892–901. [DOI] [PubMed] [Google Scholar]
  • 30.van Tilborg E, de Theije CGM, van Hal M, Wagenaar N, de Vries LS, Benders MJ, et al. Origin and dynamics of oligodendrocytes in the developing brain: Implications for perinatal white matter injury. Glia 2018, 66: 221–238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Kessaris N, Fogarty M, Iannarelli P, Grist M, Wegner M, Richardson WD. Competing waves of oligodendrocytes in the forebrain and postnatal elimination of an embryonic lineage. Nat Neurosci 2006, 9: 173–179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Tsai HH, Niu J, Munji R, Davalos D, Chang J, Zhang H, et al. Oligodendrocyte precursors migrate along vasculature in the developing nervous system. Science 2016, 351: 379–384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Masahira N, Takebayashi H, Ono K, Watanabe K, Ding L, Furusho M, et al. Olig2-positive progenitors in the embryonic spinal cord give rise not only to motoneurons and oligodendrocytes, but also to a subset of astrocytes and ependymal cells. Dev Biol 2006, 293: 358–369. [DOI] [PubMed] [Google Scholar]
  • 34.Martins-Macedo J, Lepore AC, Domingues HS, Salgado AJ, Gomes ED, Pinto L. Glial restricted precursor cells in central nervous system disorders: Current applications and future perspectives. Glia 2021, 69: 513–531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Baracskay KL, Kidd GJ, Miller RH, Trapp BD. NG2-positive cells generate A2B5-positive oligodendrocyte precursor cells. Glia 2007, 55: 1001–1010. [DOI] [PubMed] [Google Scholar]
  • 36.Levison SW, Goldman JE. Both oligodendrocytes and astrocytes develop from progenitors in the subventricular zone of postnatal rat forebrain. Neuron 1993, 10: 201–212. [DOI] [PubMed] [Google Scholar]
  • 37.Samanta J, Grund EM, Silva HM, Lafaille JJ, Fishell G, Salzer JL. Inhibition of Gli1 mobilizes endogenous neural stem cells for remyelination. Nature 2015, 526: 448–452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Butti E, Bacigaluppi M, Chaabane L, Ruffini F, Brambilla E, Berera G, et al. Neural stem cells of the subventricular zone contribute to neuroprotection of the corpus callosum after cuprizone-induced demyelination. J Neurosci 2019, 39: 5481–5492. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Xing YL, Röth PT, Stratton JAS, Chuang BHA, Danne J, Ellis SL, et al. Adult neural precursor cells from the subventricular zone contribute significantly to oligodendrocyte regeneration and remyelination. J Neurosci 2014, 34: 14128–14146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Đặng TC, Ishii Y, Nguyen V, Yamamoto S, Hamashima T, Okuno N, et al. Powerful homeostatic control of oligodendroglial lineage by PDGFRα in adult brain. Cell Rep 2019, 27: 1073–1089. [DOI] [PubMed] [Google Scholar]
  • 41.Pringle NP, Richardson WD. A singularity of PDGF alpha-receptor expression in the dorsoventral axis of the neural tube may define the origin of the oligodendrocyte lineage. Development 1993, 117: 525–533. [DOI] [PubMed] [Google Scholar]
  • 42.Schnitzer J, Schachner M. Cell type specificity of a neural cell surface antigen recognized by the monoclonal antibody A2B5. Cell Tissue Res 1982, 224: 625–636. [DOI] [PubMed] [Google Scholar]
  • 43.Stallcup WB, Arner LS, Levine JM. An antiserum against the PC12 cell line defines cell surface antigens specific for neurons and Schwann cells. J Neurosci 1983, 3: 53–68. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Huang W, Bai X, Meyer E, Scheller A. Acute brain injuries trigger microglia as an additional source of the proteoglycan NG2. Acta Neuropathol Commun 2020, 8: 146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Liu Y, Hammel G, Shi M, Cheng Z, Zivkovic S, Wang X, et al. Myelin debris stimulates NG2/CSPG4 expression in bone marrow-derived macrophages in the injured spinal cord. Front Cell Neurosci 2021, 15: 651827. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Liu YJ, Ding Y, Yin YQ, Xiao H, Hu G, Zhou JW. Cspg4high microglia contribute to microgliosis during neurodegeneration. Proc Natl Acad Sci USA 2023, 120: e2210643120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Stallcup WB, You WK, Kucharova K, Cejudo-Martin P, Yotsumoto F. NG2 proteoglycan-dependent contributions of pericytes and macrophages to brain tumor vascularization and progression. Microcirculation 2016, 23: 122–133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Fruttiger M, Karlsson L, Hall AC, Abramsson A, Calver AR, Boström H, et al. Defective oligodendrocyte development and severe hypomyelination in PDGF-a knockout mice. Development 1999, 126: 457–467. [DOI] [PubMed] [Google Scholar]
  • 49.Frost EE, Zhou Z, Krasnesky K, Armstrong RC. Initiation of oligodendrocyte progenitor cell migration by a PDGF-a activated extracellular regulated kinase (ERK) signaling pathway. Neurochem Res 2009, 34: 169–181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Zhu Q, Zhao X, Zheng K, Li H, Huang H, Zhang Z, et al. Genetic evidence that Nkx2.2 and Pdgfra are major determinants of the timing of oligodendrocyte differentiation in the developing CNS. Development 2014, 141: 548–555. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Marques S, Zeisel A, Codeluppi S, van Bruggen D, Falcão AM, Xiao L, et al. Oligodendrocyte heterogeneity in the mouse juvenile and adult central nervous system. Science 2016, 352: 1326–1329. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Rakic S, Zecevic N. Early oligodendrocyte progenitor cells in the human fetal telencephalon. Glia 2003, 41: 117–127. [DOI] [PubMed] [Google Scholar]
  • 53.Bansal R, Warrington AE, Gard AL, Ranscht B, Pfeiffer SE. Multiple and novel specificities of monoclonal antibodies O1, O4, and R-mAb used in the analysis of oligodendrocyte development. J Neurosci Res 1989, 24: 548–557. [DOI] [PubMed] [Google Scholar]
  • 54.Huang W, Bhaduri A, Velmeshev D, Wang S, Wang L, Rottkamp CA, et al. Origins and proliferative states of human oligodendrocyte precursor cells. Cell 2020, 182: 594-608.e11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Boda E, Di Maria S, Rosa P, Taylor V, Abbracchio MP, Buffo A. Early phenotypic asymmetry of sister oligodendrocyte progenitor cells after mitosis and its modulation by aging and extrinsic factors. Glia 2015, 63: 271–286. [DOI] [PubMed] [Google Scholar]
  • 56.Niu J, Yu G, Wang X, Xia W, Wang Y, Hoi KK, et al. Oligodendroglial ring finger protein Rnf43 is an essential injury-specific regulator of oligodendrocyte maturation. Neuron 2021, 109: 3104-3118.e6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Hughes EG, Kang SH, Fukaya M, Bergles DE. Oligodendrocyte progenitors balance growth with self-repulsion to achieve homeostasis in the adult brain. Nat Neurosci 2013, 16: 668–676. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Rhodes KE, Raivich G, Fawcett JW. The injury response of oligodendrocyte precursor cells is induced by platelets, macrophages and inflammation-associated cytokines. Neuroscience 2006, 140: 87–100. [DOI] [PubMed] [Google Scholar]
  • 59.von Streitberg A, Jäkel S, Eugenin von Bernhardi J, Straube C, Buggenthin F, Marr C, et al. NG2-Glia transiently overcome their homeostatic network and contribute to wound closure after brain injury. Front Cell Dev Biol 2021, 9: 662056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Nishiyama A, Chang A, Trapp BD. NG2+ glial cells: A novel glial cell population in the adult brain. J Neuropathol Exp Neurol 1999, 58: 1113–1124. [DOI] [PubMed] [Google Scholar]
  • 61.Liu R, Jia Y, Guo P, Jiang W, Bai R, Liu C. In vivo clonal analysis reveals development heterogeneity of oligodendrocyte precursor cells derived from distinct germinal zones. Adv Sci 2021, 8: 2102274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Fernandez-Castaneda A, Gaultier A. Adult oligodendrocyte progenitor cells—Multifaceted regulators of the CNS in health and disease. Brain Behav Immun 2016, 57: 1–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Crawford AH, Stockley JH, Tripathi RB, Richardson WD, Franklin RJM. Oligodendrocyte progenitors: Adult stem cells of the central nervous system? Exp Neurol 2014, 260: 50–55. [DOI] [PubMed] [Google Scholar]
  • 64.Su Y, Wang X, Yang Y, Chen L, Xia W, Hoi KK, et al. Astrocyte endfoot formation controls the termination of oligodendrocyte precursor cell perivascular migration during development. Neuron 2023, 111: 190-201.e8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Bechler ME, Swire M, Ffrench-Constant C. Intrinsic and adaptive myelination-a sequential mechanism for smart wiring in the brain. Dev Neurobiol 2018, 78: 68–79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Mount CW, Monje M. Wrapped to adapt: Experience-dependent myelination. Neuron 2017, 95: 743–756. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Williamson JM, Lyons DA. Myelin dynamics throughout life: An ever-changing landscape? Front Cell Neurosci 2018, 12: 424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Purger D, Gibson EM, Monje M. Myelin plasticity in the central nervous system. Neuropharmacology 2016, 110: 563–573. [DOI] [PubMed] [Google Scholar]
  • 69.Kaller MS, Lazari A, Blanco-Duque C, Sampaio-Baptista C, Johansen-Berg H. Myelin plasticity and behaviour-connecting the dots. Curr Opin Neurobiol 2017, 47: 86–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Levine JM, Reynolds R. Activation and proliferation of endogenous oligodendrocyte precursor cells during ethidium bromide-induced demyelination. Exp Neurol 1999, 160: 333–347. [DOI] [PubMed] [Google Scholar]
  • 71.Serwanski DR, Rasmussen AL, Brunquell CB, Perkins SS, Nishiyama A. Sequential contribution of parenchymal and neural stem cell-derived oligodendrocyte precursor cells toward remyelination. Neuroglia 2018, 1: 91–105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.David-Bercholz J, Kuo CT, Deneen B. Astrocyte and oligodendrocyte responses from the subventricular zone after injury. Front Cell Neurosci 2021, 15: 797553. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Wood PM, Bunge RP. The origin of remyelinating cells in the adult central nervous system: The role of the mature oligodendrocyte. Glia 1991, 4: 225–232. [DOI] [PubMed] [Google Scholar]
  • 74.Warrington AE, Barbarese E, Pfeiffer SE. Differential myelinogenic capacity of specific developmental stages of the oligodendrocyte lineage upon transplantation into hypomyelinating hosts. J Neurosci Res 1993, 34: 1–13. [DOI] [PubMed] [Google Scholar]
  • 75.Duncan ID, Paino C, Archer DR, Wood PM. Functional capacities of transplanted cell-sorted adult oligodendrocytes. Dev Neurosci 1992, 14: 114–122. [DOI] [PubMed] [Google Scholar]
  • 76.Yeung MSY, Djelloul M, Steiner E, Bernard S, Salehpour M, Possnert G, et al. Dynamics of oligodendrocyte generation in multiple sclerosis. Nature 2019, 566: 538–542. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Duncan ID, Radcliff AB, Heidari M, Kidd G, August BK, Wierenga LA. The adult oligodendrocyte can participate in remyelination. Proc Natl Acad Sci USA 2018, 115: E11807–E11816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Mezydlo A, Treiber N, Ullrich Gavilanes EM, Eichenseer K, Ancău M, Wens A, et al. Remyelination by surviving oligodendrocytes is inefficient in the inflamed mammalian cortex. Neuron 2023, 111: 1748-1759.e8. [DOI] [PubMed] [Google Scholar]
  • 79.Neely SA, Williamson JM, Klingseisen A, Zoupi L, Early JJ, Williams A, et al. New oligodendrocytes exhibit more abundant and accurate myelin regeneration than those that survive demyelination. Nat Neurosci 2022, 25: 415–420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Keirstead HS, Blakemore WF. Identification of post-mitotic oligodendrocytes incapable of remyelination within the demyelinated adult spinal cord. J Neuropathol Exp Neurol 1997, 56: 1191–1201. [DOI] [PubMed] [Google Scholar]
  • 81.Crawford AH, Tripathi RB, Foerster S, McKenzie I, Kougioumtzidou E, Grist M, et al. Pre-Existing Mature Oligodendrocytes Do Not Contribute to Remyelination following Toxin-Induced Spinal Cord Demyelination. Am J Pathol 2016, 186: 511–516. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Boyd A, Zhang H, Williams A. Insufficient OPC migration into demyelinated lesions is a cause of poor remyelination in MS and mouse models. Acta Neuropathol 2013, 125: 841–859. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Zuo H, Wood WM, Sherafat A, Hill RA, Lu QR, Nishiyama A. Age-dependent decline in fate switch from NG2 cells to astrocytes after Olig2 deletion. J Neurosci 2018, 38: 2359–2371. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Sawamura S, Sawada M, Ito M, Nagatsu T, Nagatsu I, Suzumura A, et al. The bipotential glial progenitor cell line can develop into both oligodendrocytes and astrocytes in the mouse forebrain. Neurosci Lett 1995, 188: 1–4. [DOI] [PubMed] [Google Scholar]
  • 85.Windrem MS, Nunes MC, Rashbaum WK, Schwartz TH, Goodman RA, McKhann G, et al. Fetal and adult human oligodendrocyte progenitor cell isolates myelinate the congenitally dysmyelinated brain. Nat Med 2004, 10: 93–97. [DOI] [PubMed] [Google Scholar]
  • 86.Uemura MT, Ihara M, Maki T, Nakagomi T, Kaji S, Uemura K, et al. Pericyte-derived bone morphogenetic protein 4 underlies white matter damage after chronic hypoperfusion. Brain Pathol 2018, 28: 521–535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Sellers DL, Maris DO, Horner PJ. Postinjury niches induce temporal shifts in progenitor fates to direct lesion repair after spinal cord injury. J Neurosci 2009, 29: 6722–6733. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Sofroniew MV. Astrocyte barriers to neurotoxic inflammation. Nat Rev Neurosci 2015, 16: 249–263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Escartin C, Galea E, Lakatos A, O’Callaghan JP, Petzold GC, Serrano-Pozo A, et al. Reactive astrocyte nomenclature, definitions, and future directions. Nat Neurosci 2021, 24: 312–325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Zawadzka M, Rivers LE, Fancy SPJ, Zhao C, Tripathi R, Jamen F, et al. CNS-resident glial progenitor/stem cells produce Schwann cells as well as oligodendrocytes during repair of CNS demyelination. Cell Stem Cell 2010, 6: 578–590. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Mount CW, Yalçın B, Cunliffe-Koehler K, Sundaresh S, Monje M. Monosynaptic tracing maps brain-wide afferent oligodendrocyte precursor cell connectivity. eLife 2019, 8: e49291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Li J, Miramontes TG, Czopka T, Monk KR. Synaptic input and Ca2+ activity in zebrafish oligodendrocyte precursor cells contribute to myelin sheath formation. Nat Neurosci 2024, 27: 219–231. [DOI] [PubMed] [Google Scholar]
  • 93.de Ceglia R, Ledonne A, Litvin DG, Lind BL, Carriero G, Latagliata EC, et al. Specialized astrocytes mediate glutamatergic gliotransmission in the CNS. Nature 2023, 622: 120–129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Badia-Soteras A, Heistek TS, Kater MSJ, Mak A, Negrean A, van den Oever MC, et al. Retraction of astrocyte leaflets from the synapse enhances fear memory. Biol Psychiatry 2023, 94: 226–238. [DOI] [PubMed] [Google Scholar]
  • 95.Kukley M, Capetillo-Zarate E, Dietrich D. Vesicular glutamate release from axons in white matter. Nat Neurosci 2007, 10: 311–320. [DOI] [PubMed] [Google Scholar]
  • 96.Xiao Y, Czopka T. Myelination-independent functions of oligodendrocyte precursor cells in health and disease. Nat Neurosci 2023, 26: 1663–1669. [DOI] [PubMed] [Google Scholar]
  • 97.Fang LP, Bai X. Oligodendrocyte precursor cells: The multitaskers in the brain. Pflugers Arch 2023, 475: 1035–1044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Buchanan J, da Costa NM, Cheadle L. Emerging roles of oligodendrocyte precursor cells in neural circuit development and remodeling. Trends Neurosci 2023, 46: 628–639. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Yi C, Verkhratsky A, Niu J. Pathological potential of oligodendrocyte precursor cells: Terra incognita. Trends Neurosci 2023, 46: 581–596. [DOI] [PubMed] [Google Scholar]
  • 100.Lepiemme F, Stoufflet J, Javier-Torrent M, Mazzucchelli G, Silva CG, Nguyen L. Oligodendrocyte precursors guide interneuron migration by unidirectional contact repulsion. Science 2022, 376: eabn6204. [DOI] [PubMed] [Google Scholar]
  • 101.Xiao Y, Petrucco L, Hoodless LJ, Portugues R, Czopka T. Oligodendrocyte precursor cells sculpt the visual system by regulating axonal remodeling. Nat Neurosci 2022, 25: 280–284. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Buchanan J, Elabbady L, Collman F, Jorstad NL, Bakken TE, Ott C, et al. Oligodendrocyte precursor cells ingest axons in the mouse neocortex. Proc Natl Acad Sci USA 2022, 119: e2202580119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Auguste YSS, Ferro A, Kahng JA, Xavier AM, Dixon JR, Vrudhula U, et al. Oligodendrocyte precursor cells engulf synapses during circuit remodeling in mice. Nat Neurosci 2022, 25: 1273–1278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Yu G, Su Y, Guo C, Yi C, Yu B, Chen H, et al. Pathological oligodendrocyte precursor cells revealed in human schizophrenic brains and trigger schizophrenia-like behaviors and synaptic defects in genetic animal model. Mol Psychiatry 2022, 27: 5154–5166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Jäkel S, Agirre E, Mendanha Falcão A, van Bruggen D, Lee KW, Knuesel I, et al. Altered human oligodendrocyte heterogeneity in multiple sclerosis. Nature 2019, 566: 543–547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Falcão AM, van Bruggen D, Marques S, Meijer M, Jäkel S, Agirre E, et al. Disease-specific oligodendrocyte lineage cells arise in multiple sclerosis. Nat Med 2018, 24: 1837–1844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Kirby BB, Takada N, Latimer AJ, Shin J, Carney TJ, Kelsh RN, et al. In vivo time-lapse imaging shows dynamic oligodendrocyte progenitor behavior during zebrafish development. Nat Neurosci 2006, 9: 1506–1511. [DOI] [PubMed] [Google Scholar]
  • 108.Marisca R, Hoche T, Agirre E, Hoodless LJ, Barkey W, Auer F, et al. Functionally distinct subgroups of oligodendrocyte precursor cells integrate neural activity and execute myelin formation. Nat Neurosci 2020, 23: 363–374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Irvine KA, Blakemore WF. A different regional response by mouse oligodendrocyte progenitor cells (OPCs) to high-dose X-irradiation has consequences for repopulating OPC-depleted normal tissue. Eur J Neuroscience 2007, 25: 417–424. [DOI] [PubMed] [Google Scholar]
  • 110.Boda E, Lorenzati M, Parolisi R, Harding B, Pallavicini G, Bonfanti L, et al. Molecular and functional heterogeneity in dorsal and ventral oligodendrocyte progenitor cells of the mouse forebrain in response to DNA damage. Nat Commun 2022, 13: 2331. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Crawford AH, Tripathi RB, Richardson WD, Franklin RJM. Developmental origin of oligodendrocyte lineage cells determines response to demyelination and susceptibility to age-associated functional decline. Cell Rep 2016, 15: 761–773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Zhu Q, Whittemore SR, Devries WH, Zhao X, Kuypers NJ, Qiu M. Dorsally-derived oligodendrocytes in the spinal cord contribute to axonal myelination during development and remyelination following focal demyelination. Glia 2011, 59: 1612–1621. [DOI] [PubMed] [Google Scholar]
  • 113.Hill RA, Patel KD, Medved J, Reiss AM, Nishiyama A. NG2 cells in white matter but not gray matter proliferate in response to PDGF. J Neurosci 2013, 33: 14558–14566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Arai K. Can oligodendrocyte precursor cells be a therapeutic target for mitigating cognitive decline in cerebrovascular disease? J Cereb Blood Flow Metab 2020, 40: 1735–1736. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Fang LP, Liu Q, Meyer E, Welle A, Huang W, Scheller A, et al. A subset of OPCs do not express Olig2 during development which can be increased in the adult by brain injuries and complex motor learning. Glia 2023, 71: 415–430. [DOI] [PubMed] [Google Scholar]
  • 116.Spitzer SO, Sitnikov S, Kamen Y, Evans KA, Kronenberg-Versteeg D, Dietmann S, et al. Oligodendrocyte progenitor cells become regionally diverse and heterogeneous with age. Neuron 2019, 101: 459-471.e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Luan W, Qi X, Liang F, Zhang X, Jin Z, Shi L, et al. Microglia impede oligodendrocyte generation in aged brain. J Inflamm Res 2021, 14: 6813–6831. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Yasuda K, Maki T, Kinoshita H, Kaji S, Toyokawa M, Nishigori R, et al. Sex-specific differences in transcriptomic profiles and cellular characteristics of oligodendrocyte precursor cells. Stem Cell Res 2020, 46: 101866. [DOI] [PubMed] [Google Scholar]
  • 119.Cerghet M, Skoff RP, Bessert D, Zhang Z, Mullins C, Ghandour MS. Proliferation and death of oligodendrocytes and myelin proteins are differentially regulated in male and female rodents. J Neurosci 2006, 26: 1439–1447. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Mizrak D, Levitin HM, Delgado AC, Crotet V, Yuan J, Chaker Z, et al. Single-cell analysis of regional differences in adult V-SVZ neural stem cell lineages. Cell Rep 2019, 26: 394-406.e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Mathys H, Davila-Velderrain J, Peng Z, Gao F, Mohammadi S, Young JZ, et al. Single-cell transcriptomic analysis of Alzheimer’s disease. Nature 2019, 570: 332–337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Choe Y, Huynh T, Pleasure SJ. Migration of oligodendrocyte progenitor cells is controlled by transforming growth factor β family proteins during corticogenesis. J Neurosci 2014, 34: 14973–14983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Merchán P, Bribián A, Sánchez-Camacho C, Lezameta M, Bovolenta P, de Castro F. Sonic hedgehog promotes the migration and proliferation of optic nerve oligodendrocyte precursors. Mol Cell Neurosci 2007, 36: 355–368. [DOI] [PubMed] [Google Scholar]
  • 124.Harlow DE, Saul KE, Komuro H, Macklin WB. Myelin proteolipid protein complexes with αv integrin and AMPA receptors in vivo and regulates AMPA-dependent oligodendrocyte progenitor cell migration through the modulation of cell-surface GluR2 expression. J Neurosci 2015, 35: 12018–12032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Tong XP, Li XY, Zhou B, Shen W, Zhang ZJ, Xu TL, et al. Ca(2+) signaling evoked by activation of Na(+) channels and Na(+)/Ca(2+) exchangers is required for GABA-induced NG2 cell migration. J Cell Biol 2009, 186: 113–128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Sanchez-Rodriguez MA, Gomez O, Esteban PF, Garcia-Ovejero D, Molina-Holgado E. The endocannabinoid 2-arachidonoylglycerol regulates oligodendrocyte progenitor cell migration. Biochem Pharmacol 2018, 157: 180–188. [DOI] [PubMed] [Google Scholar]
  • 127.Kang M, Yao Y. Laminin regulates oligodendrocyte development and myelination. Glia 2022, 70: 414–429. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Milner R, Edwards G, Streuli C, Ffrench-Constant C. A role in migration for the alpha V beta 1 integrin expressed on oligodendrocyte precursors. J Neurosci 1996, 16: 7240–7252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Simpson PB, Armstrong RC. Intracellular signals and cytoskeletal elements involved in oligodendrocyte progenitor migration. Glia 1999, 26: 22–35. [PubMed] [Google Scholar]
  • 130.Armstrong RC, Harvath L, Dubois-Dalcq ME. Type 1 astrocytes and oligodendrocyte-type 2 astrocyte glial progenitors migrate toward distinct molecules. J Neurosci Res 1990, 27: 400–407. [DOI] [PubMed] [Google Scholar]
  • 131.Zhang H, Vutskits L, Calaora V, Durbec P, Kiss JZ. A role for the polysialic acid-neural cell adhesion molecule in PDGF-induced chemotaxis of oligodendrocyte precursor cells. J Cell Sci 2004, 117: 93–103. [DOI] [PubMed] [Google Scholar]
  • 132.Milner R, Anderson HJ, Rippon RF, McKay JS, Franklin RJ, Marchionni MA, et al. Contrasting effects of mitogenic growth factors on oligodendrocyte precursor cell migration. Glia 1997, 19: 85–90. [DOI] [PubMed] [Google Scholar]
  • 133.Tsai HH, Frost E, To V, Robinson S, Ffrench-Constant C, Geertman R, et al. The chemokine receptor CXCR2 controls positioning of oligodendrocyte precursors in developing spinal cord by arresting their migration. Cell 2002, 110: 373–383. [DOI] [PubMed] [Google Scholar]
  • 134.Jarjour AA, Manitt C, Moore SW, Thompson KM, Yuh SJ, Kennedy TE. Netrin-1 is a chemorepellent for oligodendrocyte precursor cells in the embryonic spinal cord. J Neurosci 2003, 23: 3735–3744. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Tsai HH, Tessier-Lavigne M, Miller RH. Netrin 1 mediates spinal cord oligodendrocyte precursor dispersal. Development 2003, 130: 2095–2105. [DOI] [PubMed] [Google Scholar]
  • 136.Limoni G, Niquille M. Semaphorins and Plexins in central nervous system patterning: The key to it all? Curr Opin Neurobiol 2021, 66: 224–232. [DOI] [PubMed] [Google Scholar]
  • 137.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] [PubMed] [Google Scholar]
  • 138.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] [PubMed] [Google Scholar]
  • 139.Spassky N, de Castro F, Le Bras B, Heydon K, Quéraud-LeSaux F, Bloch-Gallego E, et al. Directional guidance of oligodendroglial migration by class 3 semaphorins and netrin-1. J Neurosci 2002, 22: 5992–6004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Choi JY, Jin X, Kim H, Koh S, Cho HJ, Kim BG. High mobility group box 1 as an autocrine chemoattractant for oligodendrocyte lineage cells in white matter stroke. Stroke 2023, 54: 575–586. [DOI] [PubMed] [Google Scholar]
  • 141.Zhang H, Miller RH. Density-dependent feedback inhibition of oligodendrocyte precursor expansion. J Neurosci 1996, 16: 6886–6895. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Yan H, Rivkees SA. Hypoglycemia influences oligodendrocyte development and myelin formation. Neuroreport 2006, 17: 55–59. [DOI] [PubMed] [Google Scholar]
  • 143.Niu J, Li T, Yi C, Huang N, Koulakoff A, Weng C, et al. Connexin-based channels contribute to metabolic pathways in the oligodendroglial lineage. J Cell Sci 2016, 129: 1902–1914. [DOI] [PubMed] [Google Scholar]
  • 144.Raff MC, Lillien LE. Differentiation of a bipotential glial progenitor cell: What controls the timing and the choice of developmental pathway? J Cell Sci Suppl 1988, 10: 77–83. [DOI] [PubMed] [Google Scholar]
  • 145.Neumann B, Kazanis I. Oligodendrocyte progenitor cells: The ever mitotic cells of the CNS. Front Biosci (Schol Ed) 2016, 8: 29–43. [DOI] [PubMed] [Google Scholar]
  • 146.Fressinaud C, Sarliève LL, Dalençon D, Labourdette G. Differential regulation of cerebroside sulfotransferase and 2’, 3’-cyclic nucleotide 3’-phosphodiesterase by basic fibroblast growth factor in relation to proliferation in rat oligodendrocyte cultures. J Cell Physiol 1992, 150: 34–44. [DOI] [PubMed] [Google Scholar]
  • 147.Wilson HC, Onischke C, Raine CS. Human oligodendrocyte precursor cells in vitro: Phenotypic analysis and differential response to growth factors. Glia 2003, 44: 153–165. [DOI] [PubMed] [Google Scholar]
  • 148.McTigue DM, Horner PJ, Stokes BT, Gage FH. Neurotrophin-3 and brain-derived neurotrophic factor induce oligodendrocyte proliferation and myelination of regenerating axons in the contused adult rat spinal cord. J Neurosci 1998, 18: 5354–5365. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Barres BA, Raff MC, Gaese F, Bartke I, Dechant G, Barde YA. A crucial role for neurotrophin-3 in oligodendrocyte development. Nature 1994, 367: 371–375. [DOI] [PubMed] [Google Scholar]
  • 150.Canoll PD, Musacchio JM, Hardy R, Reynolds R, Marchionni MA, Salzer JL. GGF/neuregulin is a neuronal signal that promotes the proliferation and survival and inhibits the differentiation of oligodendrocyte progenitors. Neuron 1996, 17: 229–243. [DOI] [PubMed] [Google Scholar]
  • 151.McKinnon RD, Piras G, Ida JA Jr, Dubois-Dalcq M. A role for TGF-beta in oligodendrocyte differentiation. J Cell Biol 1993, 121: 1397–1407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Grinspan JB, Edell E, Carpio DF, Beesley JS, Lavy L, Pleasure D, et al. Stage-specific effects of bone morphogenetic proteins on the oligodendrocyte lineage. J Neurobiol 2000, 43: 1–17. [PubMed] [Google Scholar]
  • 153.Iram T, Kern F, Kaur A, Myneni S, Morningstar AR, Shin H, et al. Young CSF restores oligodendrogenesis and memory in aged mice via Fgf17. Nature 2022, 605: 509–515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Küspert M, Wegner M. SomethiNG 2 talk about-Transcriptional regulation in embryonic and adult oligodendrocyte precursors. Brain Res 2016, 1638: 167–182. [DOI] [PubMed] [Google Scholar]
  • 155.Nusse R, Clevers H. Wnt/β-catenin signaling, disease, and emerging therapeutic modalities. Cell 2017, 169: 985–999. [DOI] [PubMed] [Google Scholar]
  • 156.Fancy SPJ, Baranzini SE, Zhao C, Yuk DI, Irvine KA, Kaing S, et al. Dysregulation of the Wnt pathway inhibits timely myelination and remyelination in the mammalian CNS. Genes Dev 2009, 23: 1571–1585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Sun S, Guo W, Zhang Z, Qiu M, Dai ZM. Dose-dependent regulation of oligodendrocyte specification by β-catenin signaling. Neurosci Bull 2015, 31: 271–273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Lang J, Maeda Y, Bannerman P, Xu J, Horiuchi M, Pleasure D, et al. Adenomatous polyposis coli regulates oligodendroglial development. J Neurosci 2013, 33: 3113–3130. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Azim K, Butt AM. GSK3β negatively regulates oligodendrocyte differentiation and myelination in vivo. Glia 2011, 59: 540–553. [DOI] [PubMed] [Google Scholar]
  • 160.Hennig KM, Fass DM, Zhao WN, Sheridan SD, Fu T, Erdin S, et al. WNT/β-catenin pathway and epigenetic mechanisms regulate the pitt-hopkins syndrome and schizophrenia risk gene TCF4. Mol Neuropsychiatry 2017, 3: 53–71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Dai ZM, Sun S, Wang C, Huang H, Hu X, Zhang Z, et al. Stage-specific regulation of oligodendrocyte development by Wnt/β-catenin signaling. J Neurosci 2014, 34: 8467–8473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Fancy SPJ, Harrington EP, Yuen TJ, Silbereis JC, Zhao C, Baranzini SE, et al. Axin2 as regulatory and therapeutic target in newborn brain injury and remyelination. Nat Neurosci 2011, 14: 1009–1016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Fancy SPJ, Harrington EP, Baranzini SE, Silbereis JC, Shiow LR, Yuen TJ, et al. Parallel states of pathological Wnt signaling in neonatal brain injury and colon cancer. Nat Neurosci 2014, 17: 506–512. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Hsieh J, Aimone JB, Kaspar BK, Kuwabara T, Nakashima K, Gage FH. IGF-I instructs multipotent adult neural progenitor cells to become oligodendrocytes. J Cell Biol 2004, 164: 111–122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Galvin J, Eyermann C, Colognato H. Dystroglycan modulates the ability of insulin-like growth factor-1 to promote oligodendrocyte differentiation. J Neurosci Res 2010, 88: 3295–3307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Valerio A, Ferrario M, Dreano M, Garotta G, Spano P, Pizzi M. Soluble interleukin-6 (IL-6) receptor/IL-6 fusion protein enhances in vitro differentiation of purified rat oligodendroglial lineage cells. Mol Cell Neurosci 2002, 21: 602–615. [DOI] [PubMed] [Google Scholar]
  • 167.Furusho M, Ishii A, Bansal R. Signaling by FGF receptor 2, not FGF receptor 1, regulates myelin thickness through activation of ERK1/2-MAPK, which promotes mTORC1 activity in an akt-independent manner. J Neurosci 2017, 37: 2931–2946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Wang S, Sdrulla AD, DiSibio G, Bush G, Nofziger D, Hicks C, et al. Notch receptor activation inhibits oligodendrocyte differentiation. Neuron 1998, 21: 63–75. [DOI] [PubMed] [Google Scholar]
  • 169.Xiao G, Du J, Wu H, Ge X, Xu X, Yang A, et al. Differential inhibition of Sox10 functions by notch-hes pathway. Cell Mol Neurobiol 2020, 40: 653–662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Lowery JW, Rosen V. The BMP pathway and its inhibitors in the skeleton. Physiol Rev 2018, 98: 2431–2452. [DOI] [PubMed] [Google Scholar]
  • 171.Cate HS, Sabo JK, Merlo D, Kemper D, Aumann TD, Robinson J, et al. Modulation of bone morphogenic protein signalling alters numbers of astrocytes and oligodendroglia in the subventricular zone during cuprizone-induced demyelination. J Neurochem 2010, 115: 11–22. [DOI] [PubMed] [Google Scholar]
  • 172.Samanta J, Kessler JA. Interactions between ID and OLIG proteins mediate the inhibitory effects of BMP4 on oligodendroglial differentiation. Development 2004, 131: 4131–4142. [DOI] [PubMed] [Google Scholar]
  • 173.Mabie PC, Mehler MF, Marmur R, Papavasiliou A, Song Q, Kessler JA. Bone morphogenetic proteins induce astroglial differentiation of oligodendroglial-astroglial progenitor cells. J Neurosci 1997, 17: 4112–4120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Elbaz B, Popko B. Molecular control of oligodendrocyte development. Trends Neurosci 2019, 42: 263–277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Lu QR, Sun T, Zhu Z, Ma N, Garcia M, Stiles CD, et al. Common developmental requirement for Olig function indicates a motor neuron/oligodendrocyte connection. Cell 2002, 109: 75–86. [DOI] [PubMed] [Google Scholar]
  • 176.Wegener A, Deboux C, Bachelin C, Frah M, Kerninon C, Seilhean D, et al. Gain of Olig2 function in oligodendrocyte progenitors promotes remyelination. Brain 2015, 138: 120–135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Sun T, Echelard Y, Lu R, Yuk DI, Kaing S, Stiles CD, et al. Olig bHLH proteins interact with homeodomain proteins to regulate cell fate acquisition in progenitors of the ventral neural tube. Curr Biol 2001, 11: 1413–1420. [DOI] [PubMed] [Google Scholar]
  • 178.Yu Y, Chen Y, Kim B, Wang H, Zhao C, He X, et al. Olig2 targets chromatin remodelers to enhancers to initiate oligodendrocyte differentiation. Cell 2013, 152: 248–261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Zhang K, Chen S, Yang Q, Guo S, Chen Q, Liu Z, et al. Author Correction: The Oligodendrocyte Transcription Factor 2 OLIG2 regulates transcriptional repression during myelinogenesis in rodents. Nat Commun 2022, 13: 3164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Mei F, Wang H, Liu S, Niu J, Wang L, He Y, et al. Stage-specific deletion of Olig2 conveys opposing functions on differentiation and maturation of oligodendrocytes. J Neurosci 2013, 33: 8454–8462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Meijer DH, Kane MF, Mehta S, Liu H, Harrington E, Taylor CM, et al. Separated at birth? The functional and molecular divergence of OLIG1 and OLIG2. Nat Rev Neurosci 2012, 13: 819–831. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Niu J, Mei F, Wang L, Liu S, Tian Y, Mo W, et al. Phosphorylated olig1 localizes to the cytosol of oligodendrocytes and promotes membrane expansion and maturation. Glia 2012, 60: 1427–1436. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Küspert M, Hammer A, Bösl MR, Wegner M. Olig2 regulates Sox10 expression in oligodendrocyte precursors through an evolutionary conserved distal enhancer. Nucleic Acids Res 2011, 39: 1280–1293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Stolt CC, Rehberg S, Ader M, Lommes P, Riethmacher D, Schachner M, et al. Terminal differentiation of myelin-forming oligodendrocytes depends on the transcription factor Sox10. Genes Dev 2002, 16: 165–170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Weider M, Wegener A, Schmitt C, Küspert M, Hillgärtner S, Bösl MR, et al. Elevated in vivo levels of a single transcription factor directly convert satellite glia into oligodendrocyte-like cells. PLoS Genet 2015, 11: e1005008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Bujalka H, Koenning M, Jackson S, Perreau VM, Pope B, Hay CM, et al. MYRF is a membrane-associated transcription factor that autoproteolytically cleaves to directly activate myelin genes. PLoS Biol 2013, 11: e1001625. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Emery B, Agalliu D, Cahoy JD, Watkins TA, Dugas JC, Mulinyawe SB, et al. Myelin gene regulatory factor is a critical transcriptional regulator required for CNS myelination. Cell 2009, 138: 172–185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Hornig J, Fröb F, Vogl MR, Hermans-Borgmeyer I, Tamm ER, Wegner M. The transcription factors Sox10 and Myrf define an essential regulatory network module in differentiating oligodendrocytes. PLoS Genet 2013, 9: e1003907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Liu Z, Hu X, Cai J, Liu B, Peng X, Wegner M, et al. Induction of oligodendrocyte differentiation by Olig2 and Sox10: Evidence for reciprocal interactions and dosage-dependent mechanisms. Dev Biol 2007, 302: 683–693. [DOI] [PubMed] [Google Scholar]
  • 190.Elbaz B, Aaker JD, Isaac S, Kolarzyk A, Brugarolas P, Eden A, et al. Phosphorylation state of ZFP24 controls oligodendrocyte differentiation. Cell Rep 2018, 23: 2254–2263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Howng SYB, Avila RL, Emery B, Traka M, Lin W, Watkins T, et al. ZFP191 is required by oligodendrocytes for CNS myelination. Genes Dev 2010, 24: 301–311. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Zhao C, Deng Y, Liu L, Yu K, Zhang L, Wang H, et al. Dual regulatory switch through interactions of Tcf7l2/Tcf4 with stage-specific partners propels oligodendroglial maturation. Nat Commun 2016, 7: 10883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Nakatani H, Martin E, Hassani H, Clavairoly A, Maire CL, Viadieu A, et al. Ascl1/Mash1 promotes brain oligodendrogenesis during myelination and remyelination. J Neurosci 2013, 33: 9752–9768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Kondo T, Raff M. Basic helix-loop-helix proteins and the timing of oligodendrocyte differentiation. Development 2000, 127: 2989–2998. [DOI] [PubMed] [Google Scholar]
  • 195.Liu A, Li J, Marin-Husstege M, Kageyama R, Fan Y, Gelinas C, et al. A molecular insight of Hes5-dependent inhibition of myelin gene expression: Old partners and new players. EMBO J 2006, 25: 4833–4842. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Chen XS, Zhang YH, Cai QY, Yao ZX. ID2: A negative transcription factor regulating oligodendroglia differentiation. J Neurosci Res 2012, 90: 925–932. [DOI] [PubMed] [Google Scholar]
  • 197.Kondo T, Raff M. The Id4 HLH protein and the timing of oligodendrocyte differentiation. EMBO J 2000, 19: 1998–2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Wang S, Sdrulla A, Johnson JE, Yokota Y, Barres BA. A role for the helix-loop-helix protein Id2 in the control of oligodendrocyte development. Neuron 2001, 29: 603–614. [DOI] [PubMed] [Google Scholar]
  • 199.Stolt CC, Schlierf A, Lommes P, Hillgärtner S, Werner T, Kosian T, et al. SoxD proteins influence multiple stages of oligodendrocyte development and modulate SoxE protein function. Dev Cell 2006, 11: 697–709. [DOI] [PubMed] [Google Scholar]
  • 200.Zhao C, Ma D, Zawadzka M, Fancy SPJ, Elis-Williams L, Bouvier G, et al. Sox2 Sustains Recruitment of Oligodendrocyte Progenitor Cells following CNS Demyelination and Primes Them for Differentiation during Remyelination. J Neurosci 2015, 35: 11482–11499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Zhang S, Zhu X, Gui X, Croteau C, Song L, Xu J, et al. Sox2 is essential for oligodendroglial proliferation and differentiation during postnatal brain myelination and CNS remyelination. J Neurosci 2018, 38: 1802–1820. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Zhang S, Rasai A, Wang Y, Xu J, Bannerman P, Erol D, et al. The stem cell factor Sox2 is a positive timer of oligodendrocyte development in the postnatal murine spinal cord. Mol Neurobiol 2018, 55: 9001–9015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Doi T, Ogata T, Yamauchi J, Sawada Y, Tanaka S, Nagao M. Chd7 collaborates with Sox2 to regulate activation of oligodendrocyte precursor cells after spinal cord injury. J Neurosci 2017, 37: 10290–10309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Dong F, Liu D, Jiang F, Liu Y, Wu X, Qu X, et al. Conditional deletion of Foxg1 alleviates demyelination and facilitates remyelination via the Wnt signaling pathway in cuprizone-induced demyelinated mice. Neurosci Bull 2021, 37: 15–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Cao G, Sun C, Shen H, Qu D, Shen C, Lu H. Conditional deletion of Foxg1 delayed myelination during early postnatal brain development. Int J Mol Sci 2023, 24: 13921. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Cimini A, Bernardo A, Cifone MG, Di Marzio L, Di Loreto S. TNFalpha downregulates PPARdelta expression in oligodendrocyte progenitor cells: Implications for demyelinating diseases. Glia 2003, 41: 3–14. [DOI] [PubMed] [Google Scholar]
  • 207.Huang JK, Jarjour AA, Nait Oumesmar B, Kerninon C, Williams A, Krezel W, et al. Retinoid X receptor gamma signaling accelerates CNS remyelination. Nat Neurosci 2011, 14: 45–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Koreman E, Sun X, Lu QR. Chromatin remodeling and epigenetic regulation of oligodendrocyte myelination and myelin repair. Mol Cell Neurosci 2018, 87: 18–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Zhao C, Dong C, Frah M, Deng Y, Marie C, Zhang F, et al. Dual requirement of CHD8 for chromatin landscape establishment and histone methyltransferase recruitment to promote CNS myelination and repair. Dev Cell 2018, 45: 753-768.e8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.He D, Marie C, Zhao C, Kim B, Wang J, Deng Y, et al. Chd7 cooperates with Sox10 and regulates the onset of CNS myelination and remyelination. Nat Neurosci 2016, 19: 678–689. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Huang J, Vogel G, Yu Z, Almazan G, Richard S. Type II arginine methyltransferase PRMT5 regulates gene expression of inhibitors of differentiation/DNA binding Id2 and Id4 during glial cell differentiation. J Biol Chem 2011, 286: 44424–44432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Scaglione A, Patzig J, Liang J, Frawley R, Bok J, Mela A, et al. PRMT5-mediated regulation of developmental myelination. Nat Commun 2018, 9: 2840. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Calabretta S, Vogel G, Yu Z, Choquet K, Darbelli L, Nicholson TB, et al. Loss of PRMT5 promotes PDGFRα degradation during oligodendrocyte differentiation and myelination. Dev Cell 2018, 46: 426–440. [DOI] [PubMed] [Google Scholar]
  • 214.Jablonska B, Gierdalski M, Chew LJ, Hawley T, Catron M, Lichauco A, et al. Sirt1 regulates glial progenitor proliferation and regeneration in white matter after neonatal brain injury. Nat Commun 2016, 7: 13866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Jablonska B, Adams KL, Kratimenos P, Li Z, Strickland E, Haydar TF, et al. Sirt2 promotes white matter oligodendrogenesis during development and in models of neonatal hypoxia. Nat Commun 2022, 13: 4771. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Rafalski VA, Ho PP, Brett JO, Ucar D, Dugas JC, Pollina EA, et al. Expansion of oligodendrocyte progenitor cells following SIRT1 inactivation in the adult brain. Nat Cell Biol 2013, 15: 614–624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Tiane A, Schepers M, Rombaut B, Hupperts R, Prickaerts J, Hellings N, et al. From OPC to oligodendrocyte: An epigenetic journey. Cells 2019, 8: 1236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Swiss VA, Nguyen T, Dugas J, Ibrahim A, Barres B, Androulakis IP, et al. Identification of a gene regulatory network necessary for the initiation of oligodendrocyte differentiation. PLoS ONE 2011, 6: e18088. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Moyon S, Huynh JL, Dutta D, Zhang F, Ma D, Yoo S, et al. Functional characterization of DNA methylation in the oligodendrocyte lineage. Cell Rep 2016, 15: 748–760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Tiane A, Schepers M, Riemens R, Rombaut B, Vandormael P, Somers V, et al. DNA methylation regulates the expression of the negative transcriptional regulators ID2 and ID4 during OPC differentiation. Cell Mol Life Sci 2021, 78: 6631–6644. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Zhou J, Wu YC, Xiao BJ, Guo XD, Zheng QX, Wu B. Age-related changes in the global DNA methylation profile of oligodendrocyte progenitor cells derived from rat spinal cords. Curr Med Sci 2019, 39: 67–74. [DOI] [PubMed] [Google Scholar]
  • 222.Zhang M, Wang J, Zhang K, Lu G, Liu Y, Ren K, et al. Ten-eleven translocation 1 mediated-DNA hydroxymethylation is required for myelination and remyelination in the mouse brain. Nat Commun 2021, 12: 5091. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Zhao X, He X, Han X, Yu Y, Ye F, Chen Y, et al. MicroRNA-mediated control of oligodendrocyte differentiation. Neuron 2010, 65: 612–626. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Dugas JC, Cuellar TL, Scholze A, Ason B, Ibrahim A, Emery B, et al. Dicer1 and miR-219 Are required for normal oligodendrocyte differentiation and myelination. Neuron 2010, 65: 597–611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Wang H, Moyano AL, Ma Z, Deng Y, Lin Y, Zhao C, et al. MiR-219 cooperates with miR-338 in myelination and promotes myelin repair in the CNS. Dev Cell 2017, 40: 566-582.e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Santra M, Zhang ZG, Yang J, Santra S, Santra S, Chopp M, et al. Thymosin β4 up-regulation of microRNA-146a promotes oligodendrocyte differentiation and suppression of the Toll-like proinflammatory pathway. J Biol Chem 2014, 289: 19508–19518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Zhang F, Gou Z, Zhou Y, Huang L, Shao C, Wang M, et al. MicroRNA-21-5p agomir inhibits apoptosis of oligodendrocyte precursor cell and attenuates white matter injury in neonatal rats. Brain Res Bull 2022, 189: 139–150. [DOI] [PubMed] [Google Scholar]

Articles from Neuroscience Bulletin are provided here courtesy of Springer

RESOURCES