Next Article in Journal
Expression of WNT Signaling Genes in the Dorsolateral Prefrontal Cortex in Schizophrenia
Previous Article in Journal
Impacts of Kinematic Information on Action Anticipation and the Related Neurophysiological Associations in Volleyball Experts
Previous Article in Special Issue
Clinical Potential of Transcranial Focused Ultrasound for Neurorehabilitation in Pediatric Cancer Survivors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Translating Molecular Approaches to Oligodendrocyte-Mediated Neurological Circuit Modulation

by
Jingwei Song
1,†,
Aybike Saglam
2,†,
J. Bradley Zuchero
2 and
Vivek P. Buch
2,*
1
Medical Scientist Training Program, School of Medicine, Stanford University, Stanford, CA 94305, USA
2
Department of Neurosurgery, Stanford University, Stanford, CA 94305, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Brain Sci. 2024, 14(7), 648; https://doi.org/10.3390/brainsci14070648
Submission received: 14 June 2024 / Revised: 24 June 2024 / Accepted: 25 June 2024 / Published: 27 June 2024
(This article belongs to the Special Issue Advances in Restorative Neurotherapeutic Technologies)

Abstract

:
The central nervous system (CNS) exhibits remarkable adaptability throughout life, enabled by intricate interactions between neurons and glial cells, in particular, oligodendrocytes (OLs) and oligodendrocyte precursor cells (OPCs). This adaptability is pivotal for learning and memory, with OLs and OPCs playing a crucial role in neural circuit development, synaptic modulation, and myelination dynamics. Myelination by OLs not only supports axonal conduction but also undergoes adaptive modifications in response to neuronal activity, which is vital for cognitive processing and memory functions. This review discusses how these cellular interactions and myelin dynamics are implicated in various neurocircuit diseases and disorders such as epilepsy, gliomas, and psychiatric conditions, focusing on how maladaptive changes contribute to disease pathology and influence clinical outcomes. It also covers the potential for new diagnostics and therapeutic approaches, including pharmacological strategies and emerging biomarkers in oligodendrocyte functions and myelination processes. The evidence supports a fundamental role for myelin plasticity and oligodendrocyte functionality in synchronizing neural activity and high-level cognitive functions, offering promising avenues for targeted interventions in CNS disorders.

1. Introduction

The central nervous system undergoes extensive physiological changes throughout life. Despite the majority of neurons reaching terminal differentiation early in life, the CNS has remarkable adaptability essential for learning and memory. Emerging evidence suggests that this adaptability is not strictly controlled by the neurons themselves, but that neuronal–glial interactions also influence neural circuit formation, function, and remodeling [1,2,3]. Among diverse glial populations, cells of the oligodendrocyte lineage, including oligodendrocytes (OLs) and oligodendrocyte precursor cells (OPCs), are critical for this adaptability.
OLs and OPCs are interconnected with neural circuit development, differentiation, and remodeling via molecular and structural mechanisms (Figure 1). Molecularly, OLs and OPCs express receptors for neurotransmitters and reciprocally modulate synaptic transmission. Calcium dynamics in OPCs are responsive to both behaviorally correlated neuronal activity and direct neurotransmitter exposure [4]. Reciprocally, synaptic transmission, axonal excitability, and synchronous neuronal activity are modulated by OLs and OPCs [5,6]. Structurally, OLs and OPCs modulate neuronal circuits via myelination. Myelin is formed by OLs in the CNS, whereby compact lipid-rich lamellae are wrapped around axons for proper axonal conduction [7]. Although it was long thought that myelin was relatively inert after its formation, we now know that OLs dynamically modify myelin via changes in internode length, internode spacing, and sheath thickness; these are the adaptive mechanisms essential for the modulation of conduction velocity and circuit dynamics [8,9,10,11]. These mechanisms are orchestrated by several context-dependent cell biological processes, including extracellular cues, cell adhesion dynamics, cytoskeletal organization, and exocytosis for membrane expansion [11,12,13,14,15,16,17,18,19]. Further, OLs myelinate multiple axons in the CNS, which could provide an opportunity to integrate multinodal information processing into coordinated modifications of different neural circuits at once [20,21].
Accordingly, OLs/OPCs and myelin dynamics are implicated in memory acquisition and learning [22,23,24,25]. Adaptive changes in myelin sheath length and thickness occur in response to experience or experimentally induced neuronal activity [9,24,25,26,27]. Collective molecular evidence for the role of myelin plasticity in neural circuitry has been reviewed elsewhere [28]. Clinical observations based on EEG recordings from epileptic patients also provide insights into white matter functioning in higher-level cognition. In a recent study, intracranial recordings in humans performing a cognitive task showed a significant role of white matter communicability in the preparation of goal-directed cognitive tasks [29]. Proactive control, a term that describes the cognitive control mode in which prior information prepares the brain to be sensitive to goal-directed stimuli, was shown to be dependent on white matter communicability. These findings correlate observations in the development of network controllability in the perinatal period . Further, dynamic axonal conduction velocity and associated molecular mechanisms have critical physiological roles in synaptic transmission, temporal accuracy in neuronal activity, and high-level cognition and experience-directed circuit remodeling [25,30,31,32,33]. A recent report further revealed reciprocal modulations among reward-seeking behavior, oligodendrogenesis, and myelin plasticity [34].
The molecular mechanisms of adaptive myelination are likely fundamental to CNS functions such as information processing, storage, and retrieval [35,36]. These mechanisms remain an area of active investigation. In this review, we present the current understanding of OLs, OPCs, and myelination disease and disorders of the nervous system and discuss the potential for relevant future diagnostics and therapeutics.

2. OLs, OPCs, and Myelin in Neural Circuit Disorders

2.1. Epilepsy

Myelin dynamics and OL functionality play critical roles in the modulation of neural circuitry and are key contributors to the pathophysiology of epilepsy. Epilepsy represents the aberrant hyperexcitability of neuronal networks that propagate locally or globally in the brain. The mechanisms of ictogenesis and propagation involve diverse nervous system cell types. For example, electrical activity can propagate through gap junctions mediated by glial cell types including astrocytes and oligodendrocytes [37,38,39]. In epilepsy, myelin plasticity significantly influences the orchestration of neural network dynamics by modulating the speed and synchrony of signal transmission along axons. Disruptions in OL functionality and myelin integrity are implicated in epilepsy, where maladaptive myelination triggered by seizures contributes to increased seizure burden [40]. Additionally, impaired OL function disrupts potassium buffering, which can increase seizure likelihood [41]. For example, the thalamus is a critical area for the initiation and spread of generalized epilepsy. Molecular mechanisms that transition thalamic neurons from tonic to burst firing are linked to the precision of thalamocortical circuit synchrony which is impaired in the setting of myelin damage [42,43,44].
Microglia are additionally involved in interactions among neurons, myelin, and OLs/OPCs. Microglial gain or loss of function is associated with various myelin states, suggesting a regulatory role in maintaining myelin integrity [45,46,47,48,49,50]. Microglial activation in the setting of chemotherapy-induced toxicity leads to neurotoxic astrocytic responses, a loss of OLs and OPCs, and a reduction in myelin abundance and thickness [45,51]. In these experimental models, reduced BDNF-TrkB signaling was observed. Treatment with TrkB agonists or microglia depletion altogether restored DNF expression and myelination. Dysregulated TGFβF1 signaling in mature OLs in the setting of microglial activation was suggested as a mechanistic component of hypermyelination. Single-cell transcriptomic analyses further revealed OL clusters with elevated expression of SERPINA3 and C4B, which are markers linked to uncontrolled seizures in humans and risks for other nervous system diseases and disorders [46,52,53,54,55].
Cytokine IL-33 is also involved in microglial modulation of myelination and neural circuits [56]. IL-33 mediates the fine-tuning of the excitation/inhibition (E/I) balance and synaptic plasticity via multiple mechanisms, including the stimulation of neurons in the hippocampus and the remodeling of microglial synapses in corticothalamic circuits and the spinal cord [57,58,59,60]. Conditional deletion of IL-33 in astrocytes or its cognate receptor on microglia has been linked with an increase in excitatory synapses within the corticothalamic circuit [60]. This alteration correlates with the emergence of spontaneous epileptiform activities and a heightened susceptibility to seizures by early adulthood [58,61]. In addition, IL-33-deficient mice display failed OL maturation and impaired myelin structure, with behavioral phenotypes that include anxiety and impaired social cognition [62,63].
Taken together, these findings demonstrate that white matter connectivity mediated by OLs, OPCs, myelin, microglia, and intercellular molecular mediators serves as a basis for the synchronization of neural activity. Alterations in these dynamic processes can underlie E/I imbalances, ictogenesis, and the propagation of aberrant neural activity in epilepsy.

2.2. Glioma

OLs and OPCs are involved in the structural and functional connectivity between glioma tissue and surrounding native tissue. In gliomas, malignant cells form electrically active networks via synapses among themselves and with other neurons [64,65]. Electrical connectivity has functional roles in tumor proliferation, as demonstrated by studies where optogenetic stimulation of the optic nerve pathway accelerated glioma growth in mice and a reduction in optic nerve activity hindered glioma development [66]. Similar effects were observed in the olfactory bulb, where sensory experiences and neuronal activity influenced glioma growth, with interventions reducing tumor progression [67].
Malignant neuron–glioma synapses exploit mechanisms of adaptive plasticity typically seen in healthy neuronal synapses [68]. Activity-regulated secretion of BDNF increases both the number and strength of malignant synapses via enhanced α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor (AMPAR) trafficking to the postsynaptic membrane [51]. Interestingly, in gliomas, most AMPARs are calcium-permeable compared to the typical calcium-impermeable AMPARs in normal neurons due to aberrant processing of the GluA2 mRNA [69,70,71]. Studies highlight the role of AMPAR-mediated neuron-to-glioma synapses in aiding glioma invasion and expansion in the brain, with such synapses at the invasive tumor edge promoting tumor proliferation [68,72].
Network synchrony in gliomas is facilitated by cellular coupling among glial cells, leading to synchronous calcium wave formations that support tumor proliferation. Pharmacological blockage of such synchrony led to reduced glioma proliferation in mouse models [64,68,73]. Synchronous activity is attributed to enhanced tumor survival and growth. Studies that reversed neuronal activity induced autonomous rhythmic calcium oscillations in “hub” glioma cells, which in turn influenced neural circuit adaptability and signaling pathways [74]. Disruption of the Kca3.1 channel diminished such synchrony, decreasing tumor viability and growth, thereby extending survival in animal models, suggesting the potential of HCa3.1-inhibiting drugs as a novel treatment strategy [74]. Further, thrombospondin-1 (THBS1), a calcium-signaling-associated glycoprotein and astrocyte-derived signal for synaptogenesis during development, is implicated in glioblastoma progression, particularly influencing neural circuit remodeling through its expression in various cells within the tumor microenvironment [75].
Finally, OPCs often serve as the cell of origin for high-grade gliomas [72,73,76]. The dysregulated mechanics of myelin plasticity and paracrine signaling mediated by cells of the OL lineage also influence the glioma microenvironment, which may influence response to therapy [72].

2.3. Psychiatric Disorders

Interactions between microglia and OLs/OPCs in psychiatric disorders are poorly understood, but current understandings present intriguing therapeutic implications [77]. Stressful experiences have been shown to significantly affect oligodendrogenesis and myelination, especially in the prefrontal cortex (PFC) [8]. Experimental models demonstrate that stress leads to a reduction in myelin and OL-associated proteins [78]. These dynamics are in the setting of a larger framework wherein stress and psychoactive substances affect broad classes of glial cell types, especially in the prefrontal cortex [79]. The framework for understanding myelination abnormalities in psychiatric disorders so far implicates OLs/OPCs in regional hyper/hypomyelination and a loss of axonal metabolic support, which lead to altered connectivity and axonal degeneration, which impair higher-order cognitive functions [80].
Genetic changes in psychiatric disorders notably implicate OL-, OPC-, and myelin-related genes. In a study that identified 89 abnormally regulated genes in schizophrenia, 35 of them were involved in myelination, including genes encoding for myelin proteins (MAG, MAL, MBP, PLP, MOG, CNP), growth factors and receptors (ErbB3, NRG1, BDNF), transcription factors (SOX10, Olig1, Olig2), and other myelination-associated genes (transferrin, QKI, CLDN11) [81]. For example, CNP1, a myelin sheath protein, is crucial for maintaining normal axon function and survival [81,82]. Another gene encoding a myeline component, SOX10, is located in a major susceptibility locus for schizophrenia [83]. A loss of OL-derived erb-b2 receptor tyrosine kinase 4 (EBB4) signaling can result in OL arborization and subsequent impairments in myelin and dopaminergic function, which are involved in psychiatric disorders [84]. In the setting of depression, various molecular changes are observed. Microglia-derived factors that promote OPC proliferation and remyelination are downregulated in patients with depression [85]. Single-nucleus RNAseq of the dorsolateral PFC in depression identified the greatest number of differentially expressed genes in immature clusters of OPCs [86].
Beyond impairment at the genetic level, postmortem and live imaging studies further reveal significant white matter differences in bipolar disorder (BD) [87], schizophrenia [88,89,90], major depressive disorder (MDD) [91,92,93,94], and obsessive–compulsive disorder [95,96,97,98]. These changes are located throughout diverse brain regions associated with higher-order cognitive functioning, including frontal, insular, parietal, and other cortical regions. For example, postmortem examination has revealed significant oligodendrocyte reductions in the thalamic anterior principal and centromedian nuclei in bipolar disorder patients with psychotic episodes [99]. Schizophrenia patients show decreased oligodendrocyte density in cortical and thalamic regions [100,101].
Myelin changes can be ameliorated by administering antidepressants such as desvenlafaxine, indicating a potential therapeutic pathway through the modulation of myelination processes [102,103]. The therapeutic potential of antidepressants in the context of myelin repair is gaining attention. Antidepressants like desvenlafaxine have shown promise in improving myelin and OL-related proteins in animal models, suggesting their role in myelin repair mechanisms [102,103]. This opens up new avenues for exploring antidepressant therapies not just for symptomatic management in psychiatric disorders but also for addressing underlying myelin dysregulation.
Interestingly, recent developments have implicated impaired myelination in reward-seeking neurocircuitry, which underlies the pathophysiology of substance use and feeding disorders. Reduced white matter tracts connecting nuclei relevant to the reward system were observed in neuroimaging of adolescents with anorexia nervosa [104]. Furthermore, oligodendroglial lineage cells were found to respond to dopaminergic activity elicited in the reward pathway, and experimental blockade of oligodendrogenesis reduced rodent conditioning to opioids [34].

2.4. Neurodevelopment

Myelination is intricately linked with the stages of neurological development. During development, myelination begins spontaneously in the late fetal stages and continues into early adulthood, selectively ensheathing electrically active axons [105,106,107]. Phases of myelination can proceed independently of neuronal activity or can be adaptive to neuronal activity and experience [21,31,108,109]. Myelin-driven changes in neural conduction velocity, timing, and excitability can significantly influence developmental neurological functions [110,111]. Proper myelination is crucial for achieving developmental milestones in infancy and for the maturation of complex cognitive behaviors, especially in the PFC during adolescence [105,106,112]. The causal mechanistic understanding of adaptive myelination and calcium signaling has been subject to debate [4,108,113], limited by the availability of tools and challenged by the dynamic nature of these processes. Recent advances indicate the necessity of maintaining an intact actin cytoskeleton for proper myelin sheath formation and morphology [114].
Developmental alterations in myelination can generate vulnerabilities to neurological and psychiatric disorders. In cases of depression with a history of severe child abuse, there is a notable dysregulation in the transcriptional profiles of myelin genes coupled with a reduction in myelin thickness around small-caliber axons [115]. Glutamate signaling via distinct AMPAR subunits has been identified to regulate developmental OL function, which is also found to be casually linked to cases of SCZ [116,117]. Social isolation also reduces myelin-related transcripts and causes transcriptional and ultrastructural changes in PFC OLs [102,118]. Developmental cognitive delays can be caused by neonatal hypoxia that disrupts intracellular calcium levels, a process significantly influenced by GABAergic signaling. OPCs and myelinating OLs, which express calcium-permeable channels, are especially susceptible to ischemic conditions [119]. Enzymes such as sirtulin1 respond to calcium activity by regulating OL cytoskeleton and myelin dynamics and are possibly involved in hypoxia-induced developmental cognitive delays [120].
Adaptive tuning and maintenance of the balance between excitation and inhibition (E/I balance) are crucial for normal brain development and function to refine neural communication and processing. Notably, OPCs can receive synaptic inputs from excitatory or inhibitory neurons [121]. However, the potential role or contribution of these inputs in the E/I balance is not known. Alterations in the E/I balance have been postulated to underlie the pathogenesis of several developmental neuropsychiatric disorders, including autism spectrum disorder, schizophrenia, Down’s syndrome (DS), and attention deficit/hyperactivity disorder (ADHD) [122,123,124,125,126,127].

3. Diagnostics and Therapeutic Approaches

The functional, structural, and molecular processes associated with OLs/OPCs and myelin, in conjunction with relevant microglial and neuronal biology, are extensively implicated in diverse neuropsychiatric diseases and disorders. These dynamics are candidates for future therapeutic strategies in the form of pharmacologic, biomarker-related, and neuromodulatory modalities (Figure 2).
The molecular mechanisms underpinning myelin dynamics and oligodendrocyte-mediated circuit modulation remain an area of active investigation. Beyond the aforementioned novel pharmacological possibilities, the current understanding of the molecular mechanisms underlying OLs and OPCs in disease states is insufficient to broadly permit new classes of pharmacological approaches. New pharmacological therapeutic modalities are bound to emerge as the basic biology underlying OLs/OPCs and myelination continues to mature.
Interestingly, emerging biomarkers associated with OL dysfunction are providing crucial tools for understanding disease processes. The availability of multi-omics data, including genomics, transcriptomics, and epigenomics, all possible at the single-cell level, permits unprecedented characterization of molecular and cellular states associated with disease. For example, the discovery of factors associated with oligodendrogenesis within the cerebrospinal fluid (CSF) of young mice, in addition to their fluctuating levels during aging, can be leveraged to report global neuronal myelination states [105,127]. The implications of biomarker research extend to inflammatory diseases of the central nervous system, such as multiple sclerosis and pediatric demyelinating syndromes. Antibody-based biomarkers and myelin OL glycoprotein have now been developed for pediatric demyelinating syndromes [128]. In MS, OL-specific biomarkers can be leveraged to interpret disease severity. As the field advances, the integration of multi-omics biomarker research and data processing into clinical practice will be instrumental in refining diagnostic procedures and propelling the evolution of personalized treatment paradigms in neurological disorders. The feasibility of this approach will rely on the broader progress of precision medicine and will face universal challenges in bio-data collection, analysis, regulation, and integration into healthcare systems.
Closed-loop deep brain stimulation (DBS) further offers opportunities to directly establish synchronous, or other temporally defined, activity in targeted neural circuits for functional improvement. DBS relies on intracranial electrodes to deliver electrical stimulation to anatomically defined regions. It represents a paradigmatic shift in surgical interventions of the human brain, whereby programmable electrical activity can directly modulate targeted gray and white matter regions for functional improvement. In this setting, closed-loop DBS combines real-time electrical recordings from the brain to analyze relevant disease states and gate electrical stimulation based on clinically determined parameters [129]. Such technology has been applied in the setting of episodic, fluctuating motor deficits in neuropathic pain, epilepsy, and Parkinson’s disease [130,131,132,133]. The field, to date, largely utilizes real-time recordings to determine electrical signatures of a functional brain state. However, theoretical applications extend to more nuanced restoration of neural circuit function. For example, for a hypothetical CNS nucleus whose axons innervate a destination nucleus, impaired myelination of these axons could result in impaired activity in the postsynaptic neurons (Figure 1). Closed-loops DBS, by way of real-time recording from the presynaptic cell bodies and targeted stimulation of the axons or the postsynaptic nucleus, could bypass areas of impaired myelination and directly establish electrical dynamics as a function of the recorded presynaptic activity. Such a concept would also benefit from the continued progress in high-density neural recordings and machine learning-aided interpretation of data [134].
DBS is limited in its indiscriminate stimulation of all cells and extracellular structures in the target region. In the future, direct molecular and cell-type-specific perturbations of OLs/OPCs are also possible, given recent advances in neurological gene therapy that offer the potential to genetically modify targeted cell types of the nervous system [135,136]. Chemical and genetic methods that selectively perturb non-neuronal cells of the nervous system have been shown to reverse behavioral deficits in rodent models [137,138,139,140,141]. Via minimally invasive modalities, astrocyte stimulation was directly shown to increase OLs, improve myelination, and improve behavioral outcomes in animals experiencing neuropathic pain [137]. The feasibility of targeted gene therapy toward OLs/OPCs in clinical settings is challenged by the ability to deliver transgenes to targeted cell types in the CNS and the identification of the therapeutic transgene within vector packaging limits. Possibilities in this realm will expand exponentially, with rapid progress in the fields of gene delivery to the CNS, transcriptional engineering for cell-type-specific transgene expression, and synthetic biology that provides diverse effectors for in vivo molecular modulation. For example, targeted gene delivery to OPCs may induce molecular reprogramming that promotes remyelination after axonal injury.

4. Concluding Remarks

In conclusion, the functional and structural roles of oligodendrocytes (OLs) and oligodendrocyte precursor cells (OPCs) within the central nervous system are fundamental to both the normal cognitive processes and the pathophysiology of various neuropsychiatric disorders. The dynamic interplay between myelination and neural activity not only supports cognitive functions such as learning and memory but also reveals the potential vulnerabilities where dysregulation can lead to neurological impairments. This understanding highlights the critical impact of OLs and OPCs in disease states, such as epilepsy, gliomas, and psychiatric disorders, pointing toward innovative diagnostic and therapeutic strategies that could significantly advance current treatment paradigms. By harnessing advances in molecular biology, genetics, and biomarker development, future research can enhance our ability to modulate these crucial cellular processes, potentially leading to novel and more effective treatments for a wide range of CNS disorders. This underscores the importance of continued research into the complex roles of OLs and OPCs, not only to broaden our basic scientific understanding but also to translate these findings into clinical interventions that can improve patient outcomes in neurological and psychiatric care.

Author Contributions

J.S.: writing—review and editing, writing—original draft, and conceptualization. A.S.: writing—original draft and conceptualization. J.B.Z.: writing—review and editing and conceptualization. V.P.B.: writing—review and editing, writing—original draft, and conceptualization. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

Vivek P. Buch reports consulting fees from Boston Scientific Corporation, Neuropace Inc., Insightec Ltd., Abbott Laboratories, and equity ownership in Q innovation, Inc. The remaining authors declare no conflicts of interest.

References

  1. LoTurco, J.J.; Owens, D.F.; Heath, M.J.; Davis, M.B.; Kriegstein, A.R. GABA and Glutamate Depolarize Cortical Progenitor Cells and Inhibit DNA Synthesis. Neuron 1995, 15, 1287–1298. [Google Scholar] [CrossRef] [PubMed]
  2. Luk, K.C.; Kennedy, T.E.; Sadikot, A.F. Glutamate Promotes Proliferation of Striatal Neuronal Progenitors by an NMDA Receptor-Mediated Mechanism. J. Neurosci. 2003, 23, 2239–2250. [Google Scholar] [CrossRef]
  3. Luk, K.C.; Sadikot, A.F. Glutamate and Regulation of Proliferation in the Developing Mammalian Telencephalon. Dev. Neurosci. 2004, 26, 218–228. [Google Scholar] [CrossRef] [PubMed]
  4. Lu, T.-Y.; Hanumaihgari, P.; Hsu, E.T.; Agarwal, A.; Kawaguchi, R.; Calabresi, P.A.; Bergles, D.E. Norepinephrine Modulates Calcium Dynamics in Cortical Oligodendrocyte Precursor Cells Promoting Proliferation during Arousal in Mice. Nat. Neurosci. 2023, 26, 1739–1750. [Google Scholar] [CrossRef] [PubMed]
  5. Huxley, A.F.; Stämpfli, R. Evidence for Saltatory Conduction in Peripheral Myelinated Nerve Fibres. J. Physiol. 1949, 108, 315–339. [Google Scholar] [CrossRef] [PubMed]
  6. Steadman, P.E.; Xia, F.; Ahmed, M.; Mocle, A.J.; Penning, A.R.A.; Geraghty, A.C.; Steenland, H.W.; Monje, M.; Josselyn, S.A.; Frankland, P.W. Disruption of Oligodendrogenesis Impairs Memory Consolidation in Adult Mice. Neuron 2020, 105, 150–164.e6. [Google Scholar] [CrossRef] [PubMed]
  7. Hartline, D.K.; Colman, D.R. Rapid Conduction and the Evolution of Giant Axons and Myelinated Fibers. Curr. Biol. 2007, 17, R29–R35. [Google Scholar] [CrossRef] [PubMed]
  8. Makinodan, M.; Rosen, K.M.; Ito, S.; Corfas, G. A Critical Period for Social Experience–Dependent Oligodendrocyte Maturation and Myelination. Science 2012, 337, 1357–1360. [Google Scholar] [CrossRef] [PubMed]
  9. Gibson, E.M.; Purger, D.; Mount, C.W.; Goldstein, A.K.; Lin, G.L.; Wood, L.S.; Inema, I.; Miller, S.E.; Bieri, G.; Zuchero, J.B.; et al. Neuronal Activity Promotes Oligodendrogenesis and Adaptive Myelination in the Mammalian Brain. Science 2014, 344, 1252304. [Google Scholar] [CrossRef]
  10. Osso, L.A.; Rankin, K.A.; Chan, J.R. Experience-Dependent Myelination Following Stress Is Mediated by the Neuropeptide Dynorphin. Neuron 2021, 109, 3619–3632.e5. [Google Scholar] [CrossRef]
  11. Tomassy, G.S.; Berger, D.R.; Chen, H.-H.; Kasthuri, N.; Hayworth, K.J.; Vercelli, A.; Seung, H.S.; Lichtman, J.W.; Arlotta, P. Distinct Profiles of Myelin Distribution along Single Axons of Pyramidal Neurons in the Neocortex. Science 2014, 344, 319–324.e5. [Google Scholar] [CrossRef] [PubMed]
  12. Hughes, E.G.; Orthmann-Murphy, J.L.; Langseth, A.J.; Bergles, D.E. Myelin Remodeling through Experience-Dependent Oligodendrogenesis in the Adult Somatosensory Cortex. Nat. Neurosci. 2018, 21, 696–706. [Google Scholar] [CrossRef] [PubMed]
  13. Almeida, R.G. The Rules of Attraction in Central Nervous System Myelination. Front. Cell. Neurosci. 2018, 12, 367. [Google Scholar] [CrossRef] [PubMed]
  14. Garcia, M.A.; Zuchero, J.B. Anchors Away: Glia-Neuron Adhesion Regulates Myelin Targeting and Growth. Dev. Cell 2019, 51, 659–661. [Google Scholar] [CrossRef] [PubMed]
  15. Zuchero, J.B.; Fu, M.-M.; Sloan, S.A.; Ibrahim, A.; Olson, A.; Zaremba, A.; Dugas, J.C.; Wienbar, S.; Caprariello, A.V.; Kantor, C.; et al. CNS Myelin Wrapping Is Driven by Actin Disassembly. Dev. Cell 2015, 34, 152–167. [Google Scholar] [CrossRef] [PubMed]
  16. Nawaz, S.; Sánchez, P.; Schmitt, S.; Snaidero, N.; Mitkovski, M.; Velte, C.; Brückner, B.R.; Alexopoulos, I.; Czopka, T.; Jung, S.Y.; et al. Actin Filament Turnover Drives Leading Edge Growth during Myelin Sheath Formation in the Central Nervous System. Dev. Cell 2015, 34, 139–151. [Google Scholar] [CrossRef] [PubMed]
  17. Lam, M.; Takeo, K.; Almeida, R.G.; Cooper, M.H.; Wu, K.; Iyer, M.; Kantarci, H.; Zuchero, J.B. CNS Myelination Requires VAMP2/3-Mediated Membrane Expansion in Oligodendrocytes. Nat. Commun. 2022, 13, 5583. [Google Scholar] [CrossRef] [PubMed]
  18. Pan, L.; Trimarco, A.; Zhang, A.J.; Fujimori, K.; Urade, Y.; Sun, L.O.; Taveggia, C.; Zhang, Y. Oligodendrocyte-Lineage Cell Exocytosis and L-Type Prostaglandin D Synthase Promote Oligodendrocyte Development and Myelination. Elife 2023, 12, e77441. [Google Scholar] [CrossRef] [PubMed]
  19. Snaidero, N.; Möbius, W.; Czopka, T.; Hekking, L.H.P.; Mathisen, C.; Verkleij, D.; Goebbels, S.; Edgar, J.; Merkler, D.; Lyons, D.A.; et al. Myelin Membrane Wrapping of CNS Axons by PI(3,4,5)P3-Dependent Polarized Growth at the Inner Tongue. Cell 2014, 156, 277–290. [Google Scholar] [CrossRef]
  20. Lundgaard, I.; Luzhynskaya, A.; Stockley, J.H.; Wang, Z.; Evans, K.A.; Swire, M.; Volbracht, K.; Gautier, H.O.B.; Franklin, R.J.M.; Charles, F.-C.; et al. Neuregulin and BDNF Induce a Switch to NMDA Receptor-Dependent Myelination by Oligodendrocytes. PLoS Biol. 2013, 11, e1001743. [Google Scholar] [CrossRef]
  21. Bechler, M.E.; Swire, M.; Ffrench-Constant, C. Intrinsic and Adaptive Myelination-A Sequential Mechanism for Smart Wiring in the Brain. Dev. Neurobiol. 2018, 78, 68–79. [Google Scholar] [CrossRef] [PubMed]
  22. Sampaio-Baptista, C.; Khrapitchev, A.A.; Foxley, S.; Schlagheck, T.; Scholz, J.; Jbabdi, S.; DeLuca, G.C.; Miller, K.L.; Taylor, A.; Thomas, N.; et al. Motor Skill Learning Induces Changes in White Matter Microstructure and Myelination. J. Neurosci. 2013, 33, 19499–19503. [Google Scholar] [CrossRef]
  23. McKenzie, I.A.; Ohayon, D.; Li, H.; de Faria, J.P.; Emery, B.; Tohyama, K.; Richardson, W.D. Motor Skill Learning Requires Active Central Myelination. Science 2014, 346, 318–322. [Google Scholar] [CrossRef] [PubMed]
  24. Bacmeister, C.M.; Barr, H.J.; McClain, C.R.; Thornton, M.A.; Nettles, D.; Welle, C.G.; Hughes, E.G. Motor Learning Promotes Remyelination via New and Surviving Oligodendrocytes. Nat. Neurosci. 2020, 23, 819–831. [Google Scholar] [CrossRef] [PubMed]
  25. Bacmeister, C.M.; Huang, R.; Osso, L.A.; Thornton, M.A.; Conant, L.; Chavez, A.R.; Poleg-Polsky, A.; Hughes, E.G. Motor Learning Drives Dynamic Patterns of Intermittent Myelination on Learning-Activated Axons. Nat. Neurosci. 2022, 25, 1300–1313. [Google Scholar] [CrossRef] [PubMed]
  26. Cullen, C.L.; Pepper, R.E.; Clutterbuck, M.T.; Pitman, K.A.; Oorschot, V.; Auderset, L.; Tang, A.D.; Ramm, G.; Emery, B.; Rodger, J.; et al. Periaxonal and Nodal Plasticities Modulate Action Potential Conduction in the Adult Mouse Brain. Cell Rep. 2021, 34, 108641. [Google Scholar] [CrossRef] [PubMed]
  27. Yang, S.M.; Michel, K.; Jokhi, V.; Nedivi, E.; Arlotta, P. Neuron Class–specific Responses Govern Adaptive Myelin Remodeling in the Neocortex. Science 2020, 370, eabd2109. [Google Scholar] [CrossRef] [PubMed]
  28. Kato, D.; Wake, H. Myelin Plasticity Modulates Neural Circuitry Required for Learning and Behavior. Neurosci. Res. 2021, 167, 11–16. [Google Scholar] [CrossRef] [PubMed]
  29. Buch, V.P.; Bernabei, J.M.; Ng, G.; Richardson, A.G.; Ramayya, A.; Brandon, C.; Stiso, J.; Bassett, D.S.; Lucas, T.H. “Primed to Perform”: Dynamic White Matter Graph Communicability May Drive Metastable Network Representations of Enhanced Preparatory Cognitive Control. bioRxiv 2022. [Google Scholar] [CrossRef]
  30. Arancibia-Cárcamo, I.L.; Ford, M.C.; Cossell, L.; Ishida, K.; Tohyama, K.; Attwell, D. Node of Ranvier Length as a Potential Regulator of Myelinated Axon Conduction Speed. Elife 2017, 6, e23329. [Google Scholar] [CrossRef]
  31. Munyeshyaka, M.; Fields, R.D. Oligodendroglia Are Emerging Players in Several Forms of Learning and Memory. Commun. Biol. 2022, 5, 1148. [Google Scholar] [CrossRef]
  32. Ulloa Severino, F.P.; Lawal, O.O.; Sakers, K.; Wang, S.; Kim, N.; Friedman, A.D.; Johnson, S.A.; Sriworarat, C.; Hughes, R.H.; Soderling, S.H.; et al. Training-Induced Circuit-Specific Excitatory Synaptogenesis in Mice Is Required for Effort Control. Nat. Commun. 2023, 14, 5522. [Google Scholar] [CrossRef] [PubMed]
  33. Pajevic, S.; Plenz, D.; Basser, P.J.; Fields, R.D. Oligodendrocyte-Mediated Myelin Plasticity and Its Role in Neural Synchronization. Elife 2023, 12, e81982. [Google Scholar] [CrossRef] [PubMed]
  34. Yalçın, B.; Pomrenze, M.B.; Malacon, K.; Drexler, R.; Rogers, A.E.; Shamardani, K.; Chau, I.J.; Taylor, K.R.; Ni, L.; Contreras-Esquivel, D.; et al. Myelin Plasticity in the Ventral Tegmental Area Is Required for Opioid Reward. Nature 2024, 630, 677–685. [Google Scholar] [CrossRef]
  35. Hughes, E.G.; Appel, B. The Cell Biology of CNS Myelination. Curr. Opin. Neurobiol. 2016, 39, 93–100. [Google Scholar] [CrossRef] [PubMed]
  36. Stadelmann, C.; Timmler, S.; Barrantes-Freer, A.; Simons, M. Myelin in the Central Nervous System: Structure, Function, and Pathology. Physiol. Rev. 2019, 99, 1381–1431. [Google Scholar] [CrossRef]
  37. Bedner, P.; Steinhäuser, C. Role of Impaired Astrocyte Gap Junction Coupling in Epileptogenesis. Cells 2023, 12, 1669. [Google Scholar] [CrossRef]
  38. Purnell, B.S.; Alves, M.; Boison, D. Astrocyte-Neuron Circuits in Epilepsy. Neurobiol. Dis. 2023, 179, 106058. [Google Scholar] [CrossRef]
  39. Volnova, A.; Tsytsarev, V.; Ganina, O.; Vélez-Crespo, G.E.; Alves, J.M.; Ignashchenkova, A.; Inyushin, M. The Anti-Epileptic Effects of Carbenoxolone In Vitro and In Vivo. Int. J. Mol. Sci. 2022, 23, 663. [Google Scholar] [CrossRef]
  40. Knowles, J.K.; Xu, H.; Soane, C.; Batra, A.; Saucedo, T.; Frost, E.; Tam, L.T.; Fraga, D.; Ni, L.; Villar, K.; et al. Maladaptive Myelination Promotes Generalized Epilepsy Progression. Nat. Neurosci. 2022, 25, 596–606. [Google Scholar] [CrossRef]
  41. Larson, V.A.; Mironova, Y.; Vanderpool, K.G.; Waisman, A.; Rash, J.E.; Agarwal, A.; Bergles, D.E. Oligodendrocytes Control Potassium Accumulation in White Matter and Seizure Susceptibility. Elife 2018, 7, e34829. [Google Scholar] [CrossRef]
  42. Lindquist, B.E.; Timbie, C.; Voskobiynyk, Y.; Paz, J.T. Thalamocortical Circuits in Generalized Epilepsy: Pathophysiologic Mechanisms and Therapeutic Targets. Neurobiol. Dis. 2023, 181, 106094. [Google Scholar] [CrossRef] [PubMed]
  43. Whitmire, C.J.; Liew, Y.J.; Stanley, G.B. Thalamic State Influences Timing Precision in the Thalamocortical Circuit. J. Neurophysiol. 2021, 125, 1833–1850. [Google Scholar] [CrossRef]
  44. Scott, G.; Hellyer, P.J.; Ramlackhansingh, A.F.; Brooks, D.J.; Matthews, P.M.; Sharp, D.J. Thalamic Inflammation after Brain Trauma Is Associated with Thalamo-Cortical White Matter Damage. J. Neuroinflamm. 2015, 12, 224. [Google Scholar] [CrossRef] [PubMed]
  45. Santos, E.N.; Fields, R.D. Regulation of Myelination by Microglia. Sci. Adv. 2021, 7, eabk1131. [Google Scholar] [CrossRef] [PubMed]
  46. McNamara, N.B.; Munro, D.A.D.; Bestard-Cuche, N.; Uyeda, A.; Bogie, J.F.J.; Hoffmann, A.; Holloway, R.K.; Molina-Gonzalez, I.; Askew, K.E.; Mitchell, S.; et al. Microglia Regulate Central Nervous System Myelin Growth and Integrity. Nature 2023, 613, 120–129. [Google Scholar] [CrossRef] [PubMed]
  47. Hughes, A.N.; Appel, B. Microglia Phagocytose Myelin Sheaths to Modify Developmental Myelination. Nat. Neurosci. 2020, 23, 1055–1066. [Google Scholar] [CrossRef] [PubMed]
  48. Djannatian, M.; Radha, S.; Weikert, U.; Safaiyan, S.; Wrede, C.; Deichsel, C.; Kislinger, G.; Rhomberg, A.; Ruhwedel, T.; Campbell, D.S.; et al. Myelination Generates Aberrant Ultrastructure That Is Resolved by Microglia. J. Cell Biol. 2023, 222, e202204010. [Google Scholar] [CrossRef] [PubMed]
  49. Gibson, E.M.; Nagaraja, S.; Ocampo, A.; Tam, L.T.; Wood, L.S.; Pallegar, P.N.; Greene, J.J.; Geraghty, A.C.; Goldstein, A.K.; Ni, L.; et al. Methotrexate Chemotherapy Induces Persistent Tri-Glial Dysregulation That Underlies Chemotherapy-Related Cognitive Impairment. Cell 2019, 176, 43–55.e13. [Google Scholar] [CrossRef] [PubMed]
  50. Fernández-Castañeda, A.; Lu, P.; Geraghty, A.C.; Song, E.; Lee, M.-H.; Wood, J.; O’Dea, M.R.; Dutton, S.; Shamardani, K.; Nwangwu, K.; et al. Mild Respiratory COVID Can Cause Multi-Lineage Neural Cell and Myelin Dysregulation. Cell 2022, 185, 2452–2468.e16. [Google Scholar] [CrossRef]
  51. Taylor, K.R.; Barron, T.; Hui, A.; Spitzer, A.; Yalçin, B.; Ivec, A.E.; Geraghty, A.C.; Hartmann, G.G.; Arzt, M.; Gillespie, S.M.; et al. Glioma Synapses Recruit Mechanisms of Adaptive Plasticity. Nature 2023, 623, 366–374. [Google Scholar] [CrossRef]
  52. Yang, C.; Farias, F.H.G.; Ibanez, L.; Suhy, A.; Sadler, B.; Fernandez, M.V.; Wang, F.; Bradley, J.L.; Eiffert, B.; Bahena, J.A.; et al. Genomic Atlas of the Proteome from Brain, CSF and Plasma Prioritizes Proteins Implicated in Neurological Disorders. Nat. Neurosci. 2021, 24, 1302–1312. [Google Scholar] [CrossRef]
  53. Mayilyan, K.R.; Dodds, A.W.; Boyajyan, A.S.; Soghoyan, A.F.; Sim, R.B. Complement C4B Protein in Schizophrenia. World J. Biol. Psychiatry 2008, 9, 225–230. [Google Scholar] [CrossRef]
  54. Mostafa, G.A.; Shehab, A.A. The Link of C4B Null Allele to Autism and to a Family History of Autoimmunity in Egyptian Autistic Children. J. Neuroimmunol. 2010, 223, 115–119. [Google Scholar] [CrossRef]
  55. Zhou, D.; Rudnicki, M.; Chua, G.T.; Lawrance, S.K.; Zhou, B.; Drew, J.L.; Barbar-Smiley, F.; Armstrong, T.K.; Hilt, M.E.; Birmingham, D.J.; et al. Human Complement C4B Allotypes and Deficiencies in Selected Cases with Autoimmune Diseases. Front. Immunol. 2021, 12, 739430. [Google Scholar] [CrossRef]
  56. Kenigsbuch, M.; Bost, P.; Halevi, S.; Chang, Y.; Chen, S.; Ma, Q.; Hajbi, R.; Schwikowski, B.; Bodenmiller, B.; Fu, H.; et al. A Shared Disease-Associated Oligodendrocyte Signature among Multiple CNS Pathologies. Nat. Neurosci. 2022, 25, 876–886. [Google Scholar] [CrossRef]
  57. Wang, Y.; Fu, W.-Y.; Cheung, K.; Hung, K.-W.; Chen, C.; Geng, H.; Yung, W.-H.; Qu, J.Y.; Fu, A.K.Y.; Ip, N.Y. Astrocyte-Secreted IL-33 Mediates Homeostatic Synaptic Plasticity in the Adult Hippocampus. Proc. Natl. Acad. Sci. USA 2021, 118, e2020810118. [Google Scholar] [CrossRef]
  58. Vainchtein, I.D.; Chin, G.; Cho, F.S.; Kelley, K.W.; Miller, J.G.; Chien, E.C.; Liddelow, S.A.; Nguyen, P.T.; Nakao-Inoue, H.; Dorman, L.C.; et al. Astrocyte-Derived Interleukin-33 Promotes Microglial Synapse Engulfment and Neural Circuit Development. Science 2018, 359, 1269–1273. [Google Scholar] [CrossRef]
  59. Nguyen, P.T.; Dorman, L.C.; Pan, S.; Vainchtein, I.D.; Han, R.T.; Nakao-Inoue, H.; Taloma, S.E.; Barron, J.J.; Molofsky, A.B.; Kheirbek, M.A.; et al. Microglial Remodeling of the Extracellular Matrix Promotes Synapse Plasticity. Cell 2020, 182, 388–403.e15. [Google Scholar] [CrossRef]
  60. Han, R.T.; Vainchtein, I.D.; Schlachetzki, J.C.M.; Cho, F.S.; Dorman, L.C.; Ahn, E.; Kim, D.K.; Barron, J.J.; Nakao-Inoue, H.; Molofsky, A.B.; et al. Microglial Pattern Recognition via IL-33 Promotes Synaptic Refinement in Developing Corticothalamic Circuits in Mice. J. Exp. Med. 2023, 220, e20220605. [Google Scholar] [CrossRef]
  61. Han, R.T.; Vainchtein, I.D.; Schlachetzki, J.C.M.; Cho, F.S.; Dorman, L.C.; Johung, T.; Ahn, E.; Barron, J.T.; Nakao-Inoue, H.; Joshi, A.; et al. Interleukin-33 Coordinates a Microglial Phagocytic Response and Limits Corticothalamic Excitability and Seizure Susceptibility. bioRxiv 2021. [Google Scholar] [CrossRef]
  62. Dohi, E.; Choi, E.Y.; Rose, I.V.L.; Murata, A.S.; Chow, S.; Niwa, M.; Kano, S.-I. Behavioral Changes in Mice Lacking Interleukin-33. eNeuro 2017, 4. [Google Scholar] [CrossRef]
  63. Sung, H.-Y.; Chen, W.-Y.; Huang, H.-T.; Wang, C.-Y.; Chang, S.-B.; Tzeng, S.-F. Down-Regulation of Interleukin-33 Expression in Oligodendrocyte Precursor Cells Impairs Oligodendrocyte Lineage Progression. J. Neurochem. 2019, 150, 691–708. [Google Scholar] [CrossRef]
  64. Osswald, M.; Jung, E.; Sahm, F.; Solecki, G.; Venkataramani, V.; Blaes, J.; Weil, S.; Horstmann, H.; Wiestler, B.; Syed, M.; et al. Brain Tumour Cells Interconnect to a Functional and Resistant Network. Nature 2015, 528, 93–98. [Google Scholar] [CrossRef]
  65. Winkler, F.; Wick, W. Harmful Networks in the Brain and beyond. Science 2018, 359, 1100–1101. [Google Scholar] [CrossRef]
  66. Pan, Y.; Hysinger, J.D.; Barron, T.; Schindler, N.F.; Cobb, O.; Guo, X.; Yalçın, B.; Anastasaki, C.; Mulinyawe, S.B.; Ponnuswami, A.; et al. NF1 Mutation Drives Neuronal Activity-Dependent Initiation of Optic Glioma. Nature 2021, 594, 277–282. [Google Scholar] [CrossRef]
  67. Chen, P.; Wang, W.; Liu, R.; Lyu, J.; Zhang, L.; Li, B.; Qiu, B.; Tian, A.; Jiang, W.; Ying, H.; et al. Olfactory Sensory Experience Regulates Gliomagenesis via Neuronal IGF1. Nature 2022, 606, 550–556. [Google Scholar] [CrossRef]
  68. Venkataramani, V.; Yang, Y.; Schubert, M.C.; Reyhan, E.; Tetzlaff, S.K.; Wißmann, N.; Botz, M.; Soyka, S.J.; Beretta, C.A.; Pramatarov, R.L.; et al. Glioblastoma Hijacks Neuronal Mechanisms for Brain Invasion. Cell 2022, 185, 2899–2917.e31. [Google Scholar] [CrossRef]
  69. Chen, T.-J.; Kula, B.; Nagy, B.; Barzan, R.; Gall, A.; Ehrlich, I.; Kukley, M. In Vivo Regulation of Oligodendrocyte Precursor Cell Proliferation and Differentiation by the AMPA-Receptor Subunit GluA2. Cell Rep. 2018, 25, 852–861.e7. [Google Scholar] [CrossRef]
  70. Kougioumtzidou, E.; Shimizu, T.; Hamilton, N.B.; Tohyama, K.; Sprengel, R.; Monyer, H.; Attwell, D.; Richardson, W.D. Signalling through AMPA Receptors on Oligodendrocyte Precursors Promotes Myelination by Enhancing Oligodendrocyte Survival. Elife 2017, 6, e28080. [Google Scholar] [CrossRef]
  71. Cull-Candy, S.G.; Farrant, M. Ca2+-Permeable AMPA Receptors and Their Auxiliary Subunits in Synaptic Plasticity and Disease. J. Physiol. 2021, 599, 2655–2671. [Google Scholar] [CrossRef]
  72. Taylor, K.R.; Monje, M. Neuron–oligodendroglial Interactions in Health and Malignant Disease. Nat. Rev. Neurosci. 2023, 24, 733–746. [Google Scholar] [CrossRef]
  73. Venkatesh, H.S.; Morishita, W.; Geraghty, A.C.; Silverbush, D.; Gillespie, S.M.; Arzt, M.; Tam, L.T.; Espenel, C.; Ponnuswami, A.; Ni, L.; et al. Electrical and Synaptic Integration of Glioma into Neural Circuits. Nature 2019, 573, 539–545. [Google Scholar] [CrossRef]
  74. Hausmann, D.; Hoffmann, D.C.; Venkataramani, V.; Jung, E.; Horschitz, S.; Tetzlaff, S.K.; Jabali, A.; Hai, L.; Kessler, T.; Azoŕin, D.D.; et al. Autonomous Rhythmic Activity in Glioma Networks Drives Brain Tumour Growth. Nature 2023, 613, 179–186. [Google Scholar] [CrossRef]
  75. Krishna, S.; Choudhury, A.; Keough, M.B.; Seo, K.; Ni, L.; Kakaizada, S.; Lee, A.; Aabedi, A.; Popova, G.; Lipkin, B.; et al. Glioblastoma Remodelling of Human Neural Circuits Decreases Survival. Nature 2023, 617, 599–607. [Google Scholar] [CrossRef]
  76. Venkataramani, V.; Tanev, D.I.; Strahle, C.; Studier-Fischer, A.; Fankhauser, L.; Kessler, T.; Körber, C.; Kardorff, M.; Ratliff, M.; Xie, R.; et al. Glutamatergic Synaptic Input to Glioma Cells Drives Brain Tumour Progression. Nature 2019, 573, 532–538. [Google Scholar] [CrossRef]
  77. Rahimian, R.; Perlman, K.; Canonne, C.; Mechawar, N. Targeting Microglia-Oligodendrocyte Crosstalk in Neurodegenerative and Psychiatric Disorders. Drug Discov. Today 2022, 27, 2562–2573. [Google Scholar] [CrossRef]
  78. Poggi, G.; Albiez, J.; Pryce, C.R. Effects of Chronic Social Stress on Oligodendrocyte Proliferation-Maturation and Myelin Status in Prefrontal Cortex and Amygdala in Adult Mice. Neurobiol. Stress 2022, 18, 100451. [Google Scholar] [CrossRef]
  79. Tizabi, Y.; Getachew, B.; Hauser, S.R.; Tsytsarev, V.; Manhães, A.C.; da Silva, V.D.A. Role of Glial Cells in Neuronal Function, Mood Disorders, and Drug Addiction. Brain Sci. 2024, 14, 558. [Google Scholar] [CrossRef]
  80. Nave, K.-A.; Ehrenreich, H. Myelination and Oligodendrocyte Functions in Psychiatric Diseases. JAMA Psychiatry 2014, 71, 582–584. [Google Scholar] [CrossRef]
  81. Hakak, Y.; Walker, J.R.; Li, C.; Wong, W.H.; Davis, K.L.; Buxbaum, J.D.; Haroutunian, V.; Fienberg, A.A. Genome-Wide Expression Analysis Reveals Dysregulation of Myelination-Related Genes in Chronic Schizophrenia. Proc. Natl. Acad. Sci. USA 2001, 98, 4746–4751. [Google Scholar] [CrossRef]
  82. Peirce, T.R.; Bray, N.J.; Williams, N.M.; Norton, N.; Moskvina, V.; Preece, A.; Haroutunian, V.; Buxbaum, J.D.; Owen, M.J.; O’Donovan, M.C. Convergent Evidence for 2′,3′-Cyclic Nucleotide 3′-Phosphodiesterase as a Possible Susceptibility Gene for Schizophrenia. Arch. Gen. Psychiatry 2006, 63, 18–24. [Google Scholar] [CrossRef]
  83. Iwamoto, K.; Bundo, M.; Yamada, K.; Takao, H.; Iwayama-Shigeno, Y.; Yoshikawa, T.; Kato, T. DNA Methylation Status of SOX10 Correlates with Its Downregulation and Oligodendrocyte Dysfunction in Schizophrenia. J. Neurosci. 2005, 25, 5376–5381. [Google Scholar] [CrossRef]
  84. Roy, K.; Murtie, J.C.; El-Khodor, B.F.; Edgar, N.; Sardi, S.P.; Hooks, B.M.; Benoit-Marand, M.; Chen, C.; Moore, H.; O’Donnell, P.; et al. Loss of erbB Signaling in Oligodendrocytes Alters Myelin and Dopaminergic Function, a Potential Mechanism for Neuropsychiatric Disorders. Proc. Natl. Acad. Sci. USA 2007, 104, 8131–8136. [Google Scholar] [CrossRef]
  85. Giera, S.; Luo, R.; Ying, Y.; Ackerman, S.D.; Jeong, S.-J.; Stoveken, H.M.; Folts, C.J.; Welsh, C.A.; Tall, G.G.; Stevens, B.; et al. Microglial Transglutaminase-2 Drives Myelination and Myelin Repair via GPR56/ADGRG1 in Oligodendrocyte Precursor Cells. Elife 2018, 7, e33385. [Google Scholar] [CrossRef]
  86. Nagy, C.; Maitra, M.; Tanti, A.; Suderman, M.; Théroux, J.-F.; Davoli, M.A.; Perlman, K.; Yerko, V.; Wang, Y.C.; Tripathy, S.J.; et al. Single-Nucleus Transcriptomics of the Prefrontal Cortex in Major Depressive Disorder Implicates Oligodendrocyte Precursor Cells and Excitatory Neurons. Nat. Neurosci. 2020, 23, 771–781. [Google Scholar] [CrossRef]
  87. Barysheva, M.; Jahanshad, N.; Foland-Ross, L.; Altshuler, L.L.; Thompson, P.M. White Matter Microstructural Abnormalities in Bipolar Disorder: A Whole Brain Diffusion Tensor Imaging Study. Neuroimage Clin. 2013, 2, 558–568. [Google Scholar] [CrossRef]
  88. Kubicki, M.; McCarley, R.W.; Shenton, M.E. Evidence for White Matter Abnormalities in Schizophrenia. Curr. Opin. Psychiatry 2005, 18, 121–134. [Google Scholar] [CrossRef]
  89. Lener, M.S.; Wong, E.; Tang, C.Y.; Byne, W.; Goldstein, K.E.; Blair, N.J.; Haznedar, M.M.; New, A.S.; Chemerinski, E.; Chu, K.-W.; et al. White Matter Abnormalities in Schizophrenia and Schizotypal Personality Disorder. Schizophr. Bull. 2015, 41, 300–310. [Google Scholar] [CrossRef]
  90. Stämpfli, P.; Sommer, S.; Manoliu, A.; Burrer, A.; Schmidt, A.; Herdener, M.; Seifritz, E.; Kaiser, S.; Kirschner, M. Subtle White Matter Alterations in Schizophrenia Identified with a New Measure of Fiber Density. Sci. Rep. 2019, 9, 4636. [Google Scholar] [CrossRef]
  91. Shen, X.; Reus, L.M.; Cox, S.R.; Adams, M.J.; Liewald, D.C.; Bastin, M.E.; Smith, D.J.; Deary, I.J.; Whalley, H.C.; McIntosh, A.M. Subcortical Volume and White Matter Integrity Abnormalities in Major Depressive Disorder: Findings from UK Biobank Imaging Data. Sci. Rep. 2017, 7, 5547. [Google Scholar] [CrossRef] [PubMed]
  92. He, E.; Liu, M.; Gong, S.; Fu, X.; Han, Y.; Deng, F. White Matter Alterations in Depressive Disorder. Front. Immunol. 2022, 13, 826812. [Google Scholar] [CrossRef] [PubMed]
  93. Wu, G.; Mei, B.; Hou, X.; Wang, F.; Zang, C.; Zhang, X.; Zhang, Z. White Matter Microstructure Changes in Adults with Major Depressive Disorder: Evidence from Diffusion Magnetic Resonance Imaging. BJPsych Open 2023, 9, e101. [Google Scholar] [CrossRef] [PubMed]
  94. van Velzen, L.S.; Kelly, S.; Isaev, D.; Aleman, A.; Aftanas, L.I.; Bauer, J.; Baune, B.T.; Brak, I.V.; Carballedo, A.; Connolly, C.G.; et al. White Matter Disturbances in Major Depressive Disorder: A Coordinated Analysis across 20 International Cohorts in the ENIGMA MDD Working Group. Mol. Psychiatry 2020, 25, 1511–1525. [Google Scholar] [CrossRef]
  95. Vriend, C.; de Joode, N.T.; Pouwels, P.J.W.; Liu, F.; Otaduy, M.C.G.; Pastorello, B.; Robertson, F.C.; Ipser, J.; Lee, S.; Hezel, D.M.; et al. Age of Onset of Obsessive-Compulsive Disorder Differentially Affects White Matter Microstructure. Mol. Psychiatry 2024, 29, 1033–1045. [Google Scholar] [CrossRef] [PubMed]
  96. Gan, J.; Zhong, M.; Fan, J.; Liu, W.; Niu, C.; Cai, S.; Zou, L.; Wang, Y.; Wang, Y.; Tan, C.; et al. Abnormal White Matter Structural Connectivity in Adults with Obsessive-Compulsive Disorder. Transl. Psychiatry 2017, 7, e1062. [Google Scholar] [CrossRef]
  97. Maziero, M.P.; Seitz-Holland, J.; Cho, K.I.K.; Goldenberg, J.E.; Tanamatis, T.W.; Diniz, J.B.; Cappi, C.; Alice de Mathis, M.; Otaduy, M.C.G.; da Graça Morais Martin, M.; et al. Cellular and Extracellular White Matter Abnormalities in Obsessive-Compulsive Disorder: A Diffusion Magnetic Resonance Imaging Study. Biol. Psychiatry Cogn. Neurosci. Neuroimaging 2021, 6, 983–991. [Google Scholar] [CrossRef]
  98. Azarvand Damirichi, M.; Karimi Moridani, M.; Mohammadi, S.E. Relationship between White Matter Alterations and Contamination Subgroup in Obsessive Compulsive Disorder: A Diffusion Tensor Imaging Study. Hum. Brain Mapp. 2023, 44, 3302–3310. [Google Scholar] [CrossRef] [PubMed]
  99. Byne, W.; Tatusov, A.; Yiannoulos, G.; Vong, G.S.; Marcus, S. Effects of Mental Illness and Aging in Two Thalamic Nuclei. Schizophr. Res. 2008, 106, 172–181. [Google Scholar] [CrossRef]
  100. Cassoli, J.S.; Guest, P.C.; Malchow, B.; Schmitt, A.; Falkai, P.; Martins-de-Souza, D. Disturbed Macro-Connectivity in Schizophrenia Linked to Oligodendrocyte Dysfunction: From Structural Findings to Molecules. NPJ Schizophr. 2015, 1, 15034. [Google Scholar] [CrossRef]
  101. Hof, P.R.; Haroutunian, V.; Friedrich, V.L., Jr.; Byne, W.; Buitron, C.; Perl, D.P.; Davis, K.L. Loss and Altered Spatial Distribution of Oligodendrocytes in the Superior Frontal Gyrus in Schizophrenia. Biol. Psychiatry 2003, 53, 1075–1085. [Google Scholar] [CrossRef] [PubMed]
  102. Zhou, B.; Zhu, Z.; Ransom, B.R.; Tong, X. Oligodendrocyte Lineage Cells and Depression. Mol. Psychiatry 2021, 26, 103–117. [Google Scholar] [CrossRef] [PubMed]
  103. Wang, J.; Qiao, J.; Zhang, Y.; Wang, H.; Zhu, S.; Zhang, H.; Hartle, K.; Guo, H.; Guo, W.; He, J.; et al. Desvenlafaxine Prevents White Matter Injury and Improves the Decreased Phosphorylation of the Rate-Limiting Enzyme of Cholesterol Synthesis in a Chronic Mouse Model of Depression. J. Neurochem. 2014, 131, 229–238. [Google Scholar] [CrossRef] [PubMed]
  104. Murray, S.B.; Cabeen, R.P.; Jann, K.; Tadayonnejad, R.; Strober, M.; Feusner, J.D. White Matter Microstructure in Habit and Reward Circuits in Anorexia Nervosa: Insights from a Neurite Orientation Dispersion and Density Imaging Study. Acta Psychiatr. Scand. 2023, 147, 134–144. [Google Scholar] [CrossRef] [PubMed]
  105. Russell, R.W.R. Regional Development of the Brain in Early Life. Arch. Dis. Child. 1968, 43, 388. [Google Scholar]
  106. Grydeland, H.; Vértes, P.E.; Váša, F.; Romero-Garcia, R.; Whitaker, K.; Alexander-Bloch, A.F.; Bjørnerud, A.; Patel, A.X.; Sederevicius, D.; Tamnes, C.K.; et al. Waves of Maturation and Senescence in Micro-Structural MRI Markers of Human Cortical Myelination over the Lifespan. Cereb. Cortex 2019, 29, 1369–1381. [Google Scholar] [CrossRef]
  107. de Faria, O., Jr.; Pivonkova, H.; Varga, B.; Timmler, S.; Evans, K.A.; Káradóttir, R.T. Periods of Synchronized Myelin Changes Shape Brain Function and Plasticity. Nat. Neurosci. 2021, 24, 1508–1521. [Google Scholar] [CrossRef]
  108. Battefeld, A.; Popovic, M.A.; de Vries, S.I.; Kole, M.H.P. High-Frequency Microdomain Ca2+ Transients and Waves during Early Myelin Internode Remodeling. Cell Rep. 2019, 26, 182–191.e5. [Google Scholar] [CrossRef]
  109. Pan, Y.; Monje, M. Activity Shapes Neural Circuit Form and Function: A Historical Perspective. J. Neurosci. 2020, 40, 944–954. [Google Scholar] [CrossRef]
  110. Saab, A.S.; Tzvetanova, I.D.; Nave, K.-A. The Role of Myelin and Oligodendrocytes in Axonal Energy Metabolism. Curr. Opin. Neurobiol. 2013, 23, 1065–1072. [Google Scholar] [CrossRef]
  111. Pajevic, S.; Basser, P.J.; Fields, R.D. Role of Myelin Plasticity in Oscillations and Synchrony of Neuronal Activity. Neuroscience 2014, 276, 135–147. [Google Scholar] [CrossRef]
  112. Giedd, J.N.; Blumenthal, J.; Jeffries, N.O.; Castellanos, F.X.; Liu, H.; Zijdenbos, A.; Paus, T.; Evans, A.C.; Rapoport, J.L. Brain Development during Childhood and Adolescence: A Longitudinal MRI Study. Nat. Neurosci. 1999, 2, 861–863. [Google Scholar] [CrossRef] [PubMed]
  113. Looser, Z.J.; Ravotto, L.; Jung, R.B.; Werner, H.B.; Ruhwedel, T.; Möbius, W.; Bergles, D.E.; Felipe Barros, L.; Nave, K.-A.; Weber, B.; et al. Potassium Regulates Axon-Oligodendrocyte Signaling and Metabolic Coupling in White Matter. bioRxiv 2022. [Google Scholar] [CrossRef]
  114. Iyer, M.; Kantarci, H.; Ambiel, N.; Novak, S.W.; Andrade, L.R.; Lam, M.; Münch, A.E.; Yu, X.; Khakh, B.S.; Manor, U.; et al. Oligodendrocyte Calcium Signaling Sculpts Myelin Sheath Morphology. bioRxiv 2023. [Google Scholar] [CrossRef]
  115. Lutz, P.-E.; Tanti, A.; Gasecka, A.; Barnett-Burns, S.; Kim, J.J.; Zhou, Y.; Chen, G.G.; Wakid, M.; Shaw, M.; Almeida, D.; et al. Association of a History of Child Abuse with Impaired Myelination in the Anterior Cingulate Cortex: Convergent Epigenetic, Transcriptional, and Morphological Evidence. Am. J. Psychiatry 2017, 174, 1185–1194. [Google Scholar] [CrossRef]
  116. Piller, M.; Werkman, I.L.; Brown, R.I.; Latimer, A.J.; Kucenas, S. Glutamate Signaling via the AMPAR Subunit GluR4 Regulates Oligodendrocyte Progenitor Cell Migration in the Developing Spinal Cord. J. Neurosci. 2021, 41, 5353–5371. [Google Scholar] [CrossRef] [PubMed]
  117. Liang, W.; Hou, Y.; Huang, W.; Wang, Y.; Jiang, T.; Huang, X.; Wang, Z.; Wu, F.; Zheng, J.; Zhang, J.; et al. Loss of Schizophrenia-Related miR-501-3p in Mice Impairs Sociability and Memory by Enhancing mGluR5-Mediated Glutamatergic Transmission. Sci. Adv. 2022, 8, eabn7357. [Google Scholar] [CrossRef] [PubMed]
  118. Liu, J.; Dietz, K.; DeLoyht, J.M.; Pedre, X.; Kelkar, D.; Kaur, J.; Vialou, V.; Lobo, M.K.; Dietz, D.M.; Nestler, E.J.; et al. Impaired Adult Myelination in the Prefrontal Cortex of Socially Isolated Mice. Nat. Neurosci. 2012, 15, 1621–1623. [Google Scholar] [CrossRef] [PubMed]
  119. Paez, P.M.; Lyons, D.A. Calcium Signaling in the Oligodendrocyte Lineage: Regulators and Consequences. Annu. Rev. Neurosci. 2020, 43, 163–186. [Google Scholar] [CrossRef] [PubMed]
  120. Baraban, M.; Koudelka, S.; Lyons, D.A. Ca2+ Activity Signatures of Myelin Sheath Formation and Growth In Vivo. Nat. Neurosci. 2017, 21, 19–23. [Google Scholar] [CrossRef]
  121. Bergles, D.E.; Roberts, J.D.; Somogyi, P.; Jahr, C.E. Glutamatergic Synapses on Oligodendrocyte Precursor Cells in the Hippocampus. Nature 2000, 405, 187–191. [Google Scholar] [CrossRef]
  122. Tabuchi, K.; Blundell, J.; Etherton, M.R.; Hammer, R.E.; Liu, X.; Powell, C.M.; Südhof, T.C. A Neuroligin-3 Mutation Implicated in Autism Increases Inhibitory Synaptic Transmission in Mice. Science 2007, 318, 71–76. [Google Scholar] [CrossRef]
  123. Deutsch, S.I.; Rosse, R.B.; Schwartz, B.L.; Mastropaolo, J. A Revised Excitotoxic Hypothesis of Schizophrenia: Therapeutic Implications. Clin. Neuropharmacol. 2001, 24, 43–49. [Google Scholar] [CrossRef]
  124. Kleschevnikov, A.M.; Belichenko, P.V.; Villar, A.J.; Epstein, C.J.; Malenka, R.C.; Mobley, W.C. Hippocampal Long-Term Potentiation Suppressed by Increased Inhibition in the Ts65Dn Mouse, a Genetic Model of Down Syndrome. J. Neurosci. 2004, 24, 8153–8160. [Google Scholar] [CrossRef]
  125. Gao, Y.; Hong, Y.; Huang, L.; Zheng, S.; Zhang, H.; Wang, S.; Yao, Y.; Zhao, Y.; Zhu, L.; Xu, Q.; et al. β2-Microglobulin Functions as an Endogenous NMDAR Antagonist to Impair Synaptic Function. Cell 2023, 186, 1026–1038.e20. [Google Scholar] [CrossRef]
  126. Richter, M.M.; Ehlis, A.-C.; Jacob, C.P.; Fallgatter, A.J. Cortical Excitability in Adult Patients with Attention-Deficit/hyperactivity Disorder (ADHD). Neurosci. Lett. 2007, 419, 137–141. [Google Scholar] [CrossRef]
  127. Pouget, P.; Wattiez, N.; Rivaud-Péchoux, S.; Gaymard, B. A Fragile Balance: Perturbation of GABA Mediated Circuit in Prefrontal Cortex Generates High Intraindividual Performance Variability. PLoS ONE 2009, 4, e5208. [Google Scholar] [CrossRef]
  128. Sechi, E.; Cacciaguerra, L.; Chen, J.J.; Mariotto, S.; Fadda, G.; Dinoto, A.; Lopez-Chiriboga, A.S.; Pittock, S.J.; Flanagan, E.P. Myelin Oligodendrocyte Glycoprotein Antibody-Associated Disease (MOGAD): A Review of Clinical and MRI Features, Diagnosis, and Management. Front. Neurol. 2022, 13, 885218. [Google Scholar] [CrossRef]
  129. Parastarfeizabadi, M.; Kouzani, A.Z. Advances in Closed-Loop Deep Brain Stimulation Devices. J. Neuroeng. Rehabil. 2017, 14, 79. [Google Scholar] [CrossRef]
  130. Gilron, R.; Little, S.; Perrone, R.; Wilt, R.; de Hemptinne, C.; Yaroshinsky, M.S.; Racine, C.A.; Wang, S.S.; Ostrem, J.L.; Larson, P.S.; et al. Long-Term Wireless Streaming of Neural Recordings for Circuit Discovery and Adaptive Stimulation in Individuals with Parkinson’s Disease. Nat. Biotechnol. 2021, 39, 1078–1085. [Google Scholar] [CrossRef]
  131. Sun, F.T.; Morrell, M.J. The RNS System: Responsive Cortical Stimulation for the Treatment of Refractory Partial Epilepsy. Expert Rev. Med. Devices 2014, 11, 563–572. [Google Scholar] [CrossRef] [PubMed]
  132. Shirvalkar, P.; Veuthey, T.L.; Dawes, H.E.; Chang, E.F. Closed-Loop Deep Brain Stimulation for Refractory Chronic Pain. Front. Comput. Neurosci. 2018, 12, 18. [Google Scholar] [CrossRef] [PubMed]
  133. Prosky, J.; Cagle, J.; Sellers, K.K.; Gilron, R.; de Hemptinne, C.; Schmitgen, A.; Starr, P.A.; Chang, E.F.; Shirvalkar, P. Practical Closed-Loop Strategies for Deep Brain Stimulation: Lessons From Chronic Pain. Front. Neurosci. 2021, 15, 762097. [Google Scholar] [CrossRef] [PubMed]
  134. Jun, J.J.; Steinmetz, N.A.; Siegle, J.H.; Denman, D.J.; Bauza, M.; Barbarits, B.; Lee, A.K.; Anastassiou, C.A.; Andrei, A.; Aydın, Ç.; et al. Fully Integrated Silicon Probes for High-Density Recording of Neural Activity. Nature 2017, 551, 232–236. [Google Scholar] [CrossRef] [PubMed]
  135. Hudry, E.; Vandenberghe, L.H. Therapeutic AAV Gene Transfer to the Nervous System: A Clinical Reality. Neuron 2019, 102, 263. [Google Scholar] [CrossRef] [PubMed]
  136. Song, J.; Patel, R.V.; Sharif, M.; Ashokan, A.; Michaelides, M. Chemogenetics as a Neuromodulatory Approach to Treating Neuropsychiatric Diseases and Disorders. Mol. Ther. 2022, 30, 990–1005. [Google Scholar] [CrossRef]
  137. Saika, F.; Matsuzaki, S.; Kobayashi, D.; Ideguchi, Y.; Nakamura, T.Y.; Kishioka, S.; Kiguchi, N. Chemogenetic Regulation of CX3CR1-Expressing Microglia Using Gi-DREADD Exerts Sex-Dependent Anti-Allodynic Effects in Mouse Models of Neuropathic Pain. Front. Pharmacol. 2020, 11, 925. [Google Scholar] [CrossRef]
  138. Grace, P.M.; Wang, X.; Strand, K.A.; Baratta, M.V.; Zhang, Y.; Galer, E.L.; Yin, H.; Maier, S.F.; Watkins, L.R. DREADDed Microglia in Pain: Implications for Spinal Inflammatory Signaling in Male Rats. Exp. Neurol. 2018, 304, 125–131. [Google Scholar] [CrossRef]
  139. Scofield, M.D.; Boger, H.A.; Smith, R.J.; Li, H.; Haydon, P.G.; Kalivas, P.W. Gq-DREADD Selectively Initiates Glial Glutamate Release and Inhibits Cue-Induced Cocaine Seeking. Biol. Psychiatry 2015, 78, 441–451. [Google Scholar] [CrossRef]
  140. Lu, J.; Yang, L.; Xu, Y.; Ai, L.; Chen, J.; Xiong, F.; Hu, L.; Chen, H.; Liu, J.; Yan, X.; et al. The Modulatory Effect of Motor Cortex Astrocytes on Diabetic Neuropathic Pain. J. Neurosci. 2021, 41, 5287–5302. [Google Scholar] [CrossRef]
  141. Nwachukwu, K.N.; Evans, W.A.; Sides, T.R.; Trevisani, C.P.; Davis, A.; Marshall, S.A. Chemogenetic Manipulation of Astrocytic Signaling in the Basolateral Amygdala Reduces Binge-like Alcohol Consumption in Male Mice. J. Neurosci. Res. 2021, 99, 1957–1972. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Oligodendrocytes, key to brain adaptability, produce myelin that speeds up signal propagation. Myelin’s size and morphology influence signal timing and synchronization, achieved through two interconnected processes: the formation of new myelin sheaths and the modification of existing ones. Impaired myelin dynamics result in asynchronous, spontaneous, and/or subthreshold activity in downstream neural pathways.
Figure 1. Oligodendrocytes, key to brain adaptability, produce myelin that speeds up signal propagation. Myelin’s size and morphology influence signal timing and synchronization, achieved through two interconnected processes: the formation of new myelin sheaths and the modification of existing ones. Impaired myelin dynamics result in asynchronous, spontaneous, and/or subthreshold activity in downstream neural pathways.
Brainsci 14 00648 g001
Figure 2. Summary of future diagnostic and therapeutic directions, wherein (i) clinical biomarkers including metabolomics, genomics, and neuroimaging may be used in combination with machine learning approaches to identify patients who would benefit from specific therapies, (ii) gene delivery modalities for targeted molecular perturbations, and (iii) closed-loop DBS for electrical restoration of neural connectivity implanted electrical devices.
Figure 2. Summary of future diagnostic and therapeutic directions, wherein (i) clinical biomarkers including metabolomics, genomics, and neuroimaging may be used in combination with machine learning approaches to identify patients who would benefit from specific therapies, (ii) gene delivery modalities for targeted molecular perturbations, and (iii) closed-loop DBS for electrical restoration of neural connectivity implanted electrical devices.
Brainsci 14 00648 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Song, J.; Saglam, A.; Zuchero, J.B.; Buch, V.P. Translating Molecular Approaches to Oligodendrocyte-Mediated Neurological Circuit Modulation. Brain Sci. 2024, 14, 648. https://doi.org/10.3390/brainsci14070648

AMA Style

Song J, Saglam A, Zuchero JB, Buch VP. Translating Molecular Approaches to Oligodendrocyte-Mediated Neurological Circuit Modulation. Brain Sciences. 2024; 14(7):648. https://doi.org/10.3390/brainsci14070648

Chicago/Turabian Style

Song, Jingwei, Aybike Saglam, J. Bradley Zuchero, and Vivek P. Buch. 2024. "Translating Molecular Approaches to Oligodendrocyte-Mediated Neurological Circuit Modulation" Brain Sciences 14, no. 7: 648. https://doi.org/10.3390/brainsci14070648

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

Article Metrics

Back to TopTop