Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-4rdrl Total loading time: 0 Render date: 2024-06-29T11:48:50.584Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 August 2012

Patricia Armati
Affiliation:
University of Sydney
Emily Mathey
Affiliation:
University of Sydney
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2010

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Abelson, J. F., Kwan, K. Y., O'Roak, B. J., et al. (2005). Sequence variants in SLITRK1 are associated with Tourette's syndrome. Science 310, 317–320.CrossRefGoogle ScholarPubMed
Aberg, K., Saetre, P., Lindholm, E., et al. (2006). Human QKI, a new candidate gene for schizophrenia involved in myelination. Am J Med Genet B Neuropsychiatr Genet 141, 84–90.CrossRefGoogle Scholar
Aboul-Enein, F., Bauer, J., Klein, M., et al. (2004). Selective and antigen-dependent effects of myelin degeneration on central nervous system inflammation. J Neuropathol Exp Neurol 63, 1284–1296.CrossRefGoogle ScholarPubMed
Aboul-Enein, F., Rauschka, H., Kornek, B., et al. (2003). Preferential loss of myelin-associated glycoprotein reflects hypoxia-like white matter damage in stroke and inflammatory brain diseases. J Neuropathol Exp Neurol 62, 25–33.CrossRefGoogle ScholarPubMed
Adams, R. D. and Kubik, C. S. (1952). The morbid anatomy of the demyelinating diseases. Am J Med 12, 510–546.CrossRefGoogle Scholar
Adler, C. M., Adams, J., DelBello, M. P., et al. (2006). Evidence of white matter pathology in bipolar disorder adolescents experiencing their first episode of mania: a diffusion tensor imaging study. Am J Psychiatry 163, 322–324.CrossRefGoogle ScholarPubMed
Adrian, E. K.. and Walker, B. E. (1962). Incorporation of thymidine-H3 by cells in normal and injured mouse spinal cord. J Neuropathol Exp Neurol 21, 597–609.CrossRefGoogle ScholarPubMed
Agrawal, H. C., Randle, C. L., and Agrawal, D. (1982). In vivo acylation of rat brain myelin proteolipid protein. J Biol Chem. 257, 4588–4592.Google ScholarPubMed
Aguayo, A. J., Charron, L., and Bray, G. M. (1976). Potential of Schwann cells from unmyelinated nerves to produce myelin: a quantitative ultrastructural and radiographic study. J Neurocytol 5, 565–573.CrossRefGoogle ScholarPubMed
Ainger, K., Avossa, D., Diana, A. S., Barry, C., Barbarese, E., and Carson, J. H. (1997). Transport and localization elements in myelin basic protein mRNA. J Cell Biol 138, 1077–1087.CrossRefGoogle ScholarPubMed
Ainger, K., Avossa, D., Morgan, F., et al. (1993). Transport and localization of exogenous myelin basic protein mRNA microinjected into oligodendrocytes. J Cell Biol 123, 431–441.CrossRefGoogle ScholarPubMed
Akiyama, Y., Honmou, O., Kato, T., Uede, T., Hashi, K., and Kocsis, J. D. (2001). Transplantation of clonal neural precursor cells derived from adult human brain establishes functional peripheral myelin in the rat spinal cord. Exp Neurol 167, 27–39.CrossRefGoogle ScholarPubMed
Akiyama, Y., Radtke, C., Honmou, O., and Kocsis, J. D. (2002a). Remyelination of the spinal cord following intravenous delivery of bone marrow cells. Glia 39, 229–236.CrossRefGoogle ScholarPubMed
Akiyama, Y., Radtke, C., and Kocsis, J. D. (2002b). Remyelination of the rat spinal cord by transplantation of identified bone marrow stromal cells. J Neurosci 22, 6623–6630.CrossRefGoogle ScholarPubMed
Aktas, O., Prozorovski, T., and Zipp, F. (2006). Death ligands and autoimmune demyelination. Neuroscientist 12, 305–316.CrossRefGoogle ScholarPubMed
Alberdi, E., Sanchez-Gomez, M. V., Torre, I., et al. (2006). Activation of kainate receptors sensitizes oligodendrocytes to complement attack. J Neurosci 26, 3220–3228.CrossRefGoogle ScholarPubMed
Albert, M., Antel, J. P., Bruck, W., and Stadelmann, C. (2007). Extensive cortical remyelination in patients with chronic multiple sclerosis. Brain Pathol 17, 129–138.CrossRefGoogle ScholarPubMed
Alper, G. and Schor, N. F. (2004). Toward the definition of acute disseminated encephalitis of childhood. Curr Opin Pediatr 16, 637–640.CrossRefGoogle ScholarPubMed
Ambrosini, A., Bresciani, L., Fracchia, S., Brunello, N., and Racagni, G. (1995). Metabotropic glutamate receptors negatively coupled to adenylate cyclase inhibit N-methyl-d-aspartate receptor activity and prevent neurotoxicity in mesencephalic neurons in vitro. Mol Pharmacol 47, 1057–1064.Google ScholarPubMed
Ancel, P. Y., Livinec, F., Larroque, B., et al. (2006). Cerebral palsy among very preterm children in relation to gestational age and neonatal ultrasound abnormalities: the EPIPAGE cohort study. Pediatrics 117, 828–835.CrossRefGoogle ScholarPubMed
Andre, P., Castriconi, R., Espeli, M., et al. (2004). Comparative analysis of human NK cell activation induced by NKG2D and natural cytotoxicity receptors. Eur J Immunol 34, 961–971.CrossRefGoogle ScholarPubMed
Anitei, M., Cowan, A. E., Pfeiffer, S. E., and Bansal, R. (2009). Role for Rab3a in oligodendrocyte morphological differentiation. J Neurosci Res 87, 342–352.CrossRefGoogle ScholarPubMed
Antel, J. P., and Bar-Or, A. (2003). Do myelin-directed antibodies predict multiple sclerosis? N Engl J Med 349, 107–109.CrossRefGoogle ScholarPubMed
Antel, J. P., McCrea, E., Ladiwala, U., Qin, Y. F., and Becher, B. (1998). Non-MHC-restricted cell-mediated lysis of human oligodendrocytes in vitro: relation with CD56 expression. J Immunol 160, 1606–1611.Google ScholarPubMed
Anton, E. S., Hadjiargyrou, M., Patterson, P. H., and Matthew, W. D. (1995). CD9 plays a role in Schwann cell migration in vitro. J Neurosci 15, 584–595.CrossRefGoogle Scholar
Antony, J. M., Marle, G., Opii, W., et al. (2004). Human endogenous retrovirus glycoprotein-mediated induction of redox reactants causes oligodendrocyte death and demyelination. Nat Neurosci 7, 1088–1095.CrossRefGoogle ScholarPubMed
Aquino, J. B., Hjerling-Leffler, J., Koltzenburg, M., Edlund, T., Villar, M. J., and Ernfors, P. (2006). In vitro and in vivo differentiation of boundary cap neural crest stem cells into mature Schwann cells. Exp Neurol 198, 438–449.CrossRefGoogle ScholarPubMed
Ara, J., Przedborski, S., Naini, A. B., et al. (1998). Inactivation of tyrosine hydroxylase by nitration following exposure to peroxynitrite and 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP). Proc Natl Acad Sci U S A 95, 7659–7663.CrossRefGoogle Scholar
Araki, T. and Milbrandt, J. (1996). Ninjurin, a novel adhesion molecule, is induced by nerve injury and promotes axonal growth. Neuron 17, 353–361.CrossRefGoogle ScholarPubMed
Arbuthnott, E. R., Boyd, I. A., and Kalu, K. U. (1980). Ultrastructural dimensions of myelinated peripheral nerve fibres in the cat and their relation to conduction velocity. J Physiol (Lond) 308, 125–157.CrossRefGoogle ScholarPubMed
Archelos, J. J., Previtali, S. C., and Hartung, H. P. (1999). The role of integrins in immune-mediated diseases of the nervous system. Trends Neurosci 22, 30–38.CrossRefGoogle ScholarPubMed
Arnett, H. A., Fancy, S. P., Alberta, J. A., et al. (2004). bHLH transcription factor Olig1 is required to repair demyelinated lesions in the CNS. Science 306, 2111–2115.CrossRefGoogle ScholarPubMed
Arquint, M., Roder, J., Chia, L.-S., et al. (1987). Molecular cloning and primary structure of myelin-associated glycoproteins. Proc Natl Acad Sci USA 84, 600–604.CrossRefGoogle Scholar
Arriza, J. L., Fairman, W. A., Wadiche, J. I., Murdoch, G. H., Kavanaugh, M. P., and Amara, S. G. (1994). Functional comparisons of three glutamate transporter subtypes cloned from human motor cortex. J Neurosci 14, 5559–5569.CrossRefGoogle ScholarPubMed
Arundine, M. and Tymianski, M. (2004). Molecular mechanisms of glutamate-dependent neurodegeneration in ischemia and traumatic brain injury. Cell Mol Life Sci 61, 657–668.CrossRefGoogle ScholarPubMed
Atlante, A., Calissano, P., Bobba, A., Giannattasio, S., Marra, E., and Passarella, S. (2001). Glutamate neurotoxicity, oxidative stress and mitochondria. FEBS Lett 497, 1–5.CrossRefGoogle ScholarPubMed
Azari, M. F., Profyris, C., Karnezis, T., et al. (2006). Leukemia inhibitory factor arrests oligodendrocyte death and demyelination in spinal cord injury. J Neuropathol Exp Neurol 65, 914–929.CrossRefGoogle ScholarPubMed
Babbe, H., Roers, A., Waisman, A., et al. (2000). Clonal expansions of CD8(+) T cells dominate the T cell infiltrate in active multiple sclerosis lesions as shown by micromanipulation and single cell polymerase chain reaction. J Exp Med 192, 393–404.CrossRefGoogle Scholar
Bachelin, C., Lachapelle, F., Girard, C., et al. (2005). Efficient myelin repair in the macaque spinal cord by autologous grafts of Schwann cells. Brain 128, 540–549.CrossRefGoogle ScholarPubMed
Back, S. A., Craig, A., Kayton, R. J., et al. (2007). Hypoxia-ischemia preferentially triggers glutamate depletion from oligodendroglia and axons in perinatal cerebral white matter. J Cereb Blood Flow Metab 27, 334–347.CrossRefGoogle ScholarPubMed
Back, S. A., Gan, X., Li, Y., Rosenberg, P. A., and Volpe, J. J. (1998). Maturation-dependent vulnerability of oligodendrocytes to oxidative stress-induced death caused by glutathione depletion. J Neurosci 18, 6241–6253.CrossRefGoogle ScholarPubMed
Back, S. A., Han, B. H., Luo, N. L., et al. (2002a). Selective vulnerability of late oligodendrocyte progenitors to hypoxia-ischemia. J Neurosci 22, 455–463.CrossRefGoogle ScholarPubMed
Back, S. A., Luo, N. L., Borenstein, N. S., Levine, J. M., Volpe, J. J., and Kinney, H. C. (2001). Late oligodendrocyte progenitors coincide with the developmental window of vulnerability for human perinatal white matter injury. J Neurosci 21, 1302–1312.CrossRefGoogle ScholarPubMed
Back, S. A., Luo, N. L., Borenstein, N. S., Volpe, J. J., and Kinney, H. C. (2002b). Arrested oligodendrocyte lineage progression during human cerebral white matter development: dissociation between the timing of progenitor differentiation and myelinogenesis. J Neuropathol Exp Neurol 61, 197–211.CrossRefGoogle ScholarPubMed
Back, S. A., Tuohy, T. M., Chen, H., et al. (2005). Hyaluronan accumulates in demyelinated lesions and inhibits oligodendrocyte progenitor maturation. Nat Med 11, 966–972.CrossRefGoogle ScholarPubMed
Badea, T. C., Niculescu, F. I., Soane, L., Shin, M. L., and Rus, H. (1998). Molecular cloning and characterization of RGC-32, a novel gene induced by complement activation in oligodendrocytes. J Biol Chem 273, 26977–26981.CrossRefGoogle ScholarPubMed
Baechner, D., Liehr, T., Hameister, H., et al. (1995). Widespread expression of the peripheral myelin protein-22 gene (PMP22) in neural and non-neural tissues during murine development. J Neurosci Res 42, 733–741.CrossRefGoogle ScholarPubMed
Baerwald, K. D. and Popko, B. (1998). Developing and mature oligodendrocytes respond differently to the immune cytokine interferon-gamma. J Neurosci Res 52, 230–239.3.0.CO;2-B>CrossRefGoogle ScholarPubMed
Bagasra, O., Michaels, F. H., Zheng, Y. M., et al. (1995). Activation of the inducible form of nitric oxide synthase in the brains of patients with multiple sclerosis. Proc Natl Acad Sci USA 92, 12041–12045.CrossRefGoogle ScholarPubMed
Bai, L., Caplan, A., Lennon, D., and Miller, R. H. (2007). Human mesenchymal stem cells signals regulate neural stem cell fate. Neurochem Res 32, 353–362.CrossRefGoogle ScholarPubMed
Bakiri, Y., Burzomato, V., Frugier, G., Hamilton, N. B., Karadottir, R., and Attwell, D. (2009). Glutamatergic signaling in the brain's white matter. Neuroscience 158, 266–274.CrossRefGoogle ScholarPubMed
Bakiri, Y., Hamilton, N. B., Karadottir, R., and Attwell, D. (2008). Testing NMDA receptor block as a therapeutic strategy for reducing ischaemic damage to CNS white matter. Glia 56, 233–240.CrossRefGoogle Scholar
Bal-Price, A. and Brown, G. C. (2001). Inflammatory neurodegeneration mediated by nitric oxide from activated glia-inhibiting neuronal respiration, causing glutamate release and excitotoxicity. J Neurosci 21, 6480–6491.CrossRefGoogle ScholarPubMed
Balabanov, R., Strand, K., Goswami, R., et al. (2007). Interferon-gamma-oligodendrocyte interactions in the regulation of experimental autoimmune encephalomyelitis. J Neurosci 27, 2013–2024.CrossRefGoogle ScholarPubMed
Balentine, J. D. (1978a). Pathology of experimental spinal cord trauma. I. The necrotic lesion as a function of vascular injury. Lab Invest 39, 236–253.Google ScholarPubMed
Balentine, J. D. (1978b). Pathology of experimental spinal cord trauma. II. Ultrastructure of axons and myelin. Lab Invest 39, 254–266.Google ScholarPubMed
Balentine, J. D., and Spector, M. (1977). Calcification of axons in experimental spinal cord trauma. Ann Neurol 2, 520–523.CrossRefGoogle ScholarPubMed
Balice-Gordon, R. J., Bone, L. J., and Scherer, S. S. (1998). Functional gap junctions in the Schwann cell myelin sheath. J Cell Biol 142, 1095–1104.CrossRefGoogle ScholarPubMed
Bandtlow, C., Zachleder, T., and Schwab, M. E. (1990). Oligodendrocytes arrest neurite growth by contact inhibition. J Neurosci 10, 3837–3848.CrossRefGoogle ScholarPubMed
Banker, B. Q., and Larroche, J. C. (1962). Periventricular leukomalacia of infancy. A form of neonatal anoxic encephalopathy. Arch Neurol 7, 386–410.CrossRefGoogle ScholarPubMed
Bansal, R. and Pfeiffer, S. E. (1992). Novel stage in the oligodendrocyte lineage defined by reactivity of progenitors with R-mAb prior to O1 anti-galactocerebroside. J Neurosci Res 32, 309–316.CrossRefGoogle ScholarPubMed
Bansal, R. and Pfeiffer, S. E. (1997). FGF-2 converts mature oligodendrocytes to a novel phenotype. J Neurosci Res 50, 215–228.3.0.CO;2-7>CrossRefGoogle ScholarPubMed
Baracskay, K. L., Kidd, G. J., Miller, R. H., and Trapp, B. D. (2007). NG2-positive cells generate A2B5-positive oligodendrocyte precursor cells. Glia 55, 1001–1010.CrossRefGoogle ScholarPubMed
Barbarese, E., Carson, J. H., and Braun, P. E. (1978). Accumulation of the four myelin basic proteins in mouse brain during development. J Neurochem 31, 779–782.CrossRefGoogle ScholarPubMed
Barbarese, E., Koppel, D. E., Deutscher, M. P., et al. (1995). Protein translation components are colocalized in granules in oligodendrocytes. J Cell Sci 108, 2781–2790.Google ScholarPubMed
Barkhof, F., Bruck, W., Groot, C. J., et al. (2003). Remyelinated lesions in multiple sclerosis: magnetic resonance image appearance. Arch Neurol 60, 1073–1081.CrossRefGoogle ScholarPubMed
Barnett, M. H. and Prineas, J. W. (2004). Relapsing and remitting multiple sclerosis: pathology of the newly forming lesion. Ann Neurol 55, 458–468.CrossRefGoogle ScholarPubMed
Barradas, P. C., Ferraz, A. S., Ferreira, A. A., Daumas, R. P., and Moura, E. G. (2000). 2′3 ′-Cyclic nucleotide 3′-phosphodiesterase immunohistochemistry shows an impairment on myelin compaction in hypothyroid rats. Int J Dev Neurosci 18, 887–892.CrossRefGoogle Scholar
Barres, B. A., Koroshetz, W. J., Swartz, K. J., Chun, L. L. Y., and Corey, D. P. (1990). Ion channel expression by white matter glia: the O2A glial progenitor cell. Neuron 4, 507–524.CrossRefGoogle Scholar
Bartel, D. P. (2004). MicroRNAs: genomics, biogenesis, mechanism, and function. Cell 116, 281–297.CrossRefGoogle ScholarPubMed
Barton, W. A., Liu, B. P., Tzvetkova, D., et al. (2003). Structure and axon outgrowth inhibitor binding of the Nogo-66 receptor and related proteins. EMBO J 22, 3291–3302.CrossRefGoogle ScholarPubMed
Bartsch, S., Montag, D., Schachner, M., and Bartsch, U. (1997). Increased number of unmyelinated axons in optic nerves of adult mice deficient in the myelin-associated glycoprotein (MAG). Brain Res 762, 231–234.CrossRefGoogle Scholar
Bartsch, U. (2003). Neural CAMS and their role in the development and organization of myelin sheaths. Front Biosci 8, d477–d490.CrossRefGoogle ScholarPubMed
Bartsch, U., Bandtlow, C. E., Schnell, L., et al. (1995a). Lack of evidence that myelin-associated glycoprotein is a major inhibitor of axonal regeneration in the CNS. Neuron 15, 1375–1381.CrossRefGoogle ScholarPubMed
Bartsch, U., Kirchhoff, F., and Schachner, M. (1989). Immunohistological localization of the adhesion molecules L1, N-CAM, and MAG in the developing and adult optic nerve of mice. J Comp Neurol 284, 451–462.CrossRefGoogle Scholar
Bartsch, U., Montag, D., Bartsch, S., and Schachner, M. (1995b). Multiply myelinated axons in the optic nerve of mice deficient for the myelin-associated glycoprotein. Glia 14, 115–122.CrossRefGoogle ScholarPubMed
Bartsch, U., Pesheva, P., Raff, M., and Schachner, M. (1993). Expression of janusin (J1–160/180) in the retina and optic nerve of the developing and adult mouse. Glia 9, 57–69.CrossRefGoogle ScholarPubMed
Bartzokis, G. (2005). Brain myelination in prevalent neuropsychiatric developmental disorders: primary and comorbid addiction. Adolesc Psychiatry 29, 55–96.Google ScholarPubMed
Bartzokis, G. (2007). Acetylcholinesterase inhibitors may improve myelin integrity. Biol Psychiatry 62, 294–301.CrossRefGoogle ScholarPubMed
Bartzokis, G., Lu, P. H., Geschwind, D. H., Edwards, N., Mintz, J., and Cummings, J. L. (2006). Apolipoprotein E genotype and age-related myelin breakdown in healthy individuals: implications for cognitive decline and dementia. Arch Gen Psychiatry 63, 63–72.CrossRefGoogle ScholarPubMed
Baumann, N. and Pham-Dinh, D. (2001). Biology of oligodendrocyte and myelin in the mammalian central nervous system. Physiol Rev 81, 871–927.CrossRefGoogle ScholarPubMed
Beattie, M. S. (2004). Inflammation and apoptosis: linked therapeutic targets in spinal cord injury. Trends Mol Med 10, 580–583.CrossRefGoogle ScholarPubMed
Bechmann, I., Galea, I., and Perry, V. H. (2007). What is the blood-brain barrier (not)? Trends Immunol 28, 5–11.CrossRefGoogle Scholar
Becker, T., Anliker, B., Becker, C. G., et al. (2000). Tenascin-R inhibits regrowth of optic fibers in vitro and persists in the optic nerve of mice after injury. Glia 29, 330–346.3.0.CO;2-L>CrossRefGoogle ScholarPubMed
Behrens, T. E., Johansen-Berg, H., Woolrich, M. W., et al. (2003). Non-invasive mapping of connections between human thalamus and cortex using diffusion imaging. Nat Neurosci 6, 750–757.CrossRefGoogle ScholarPubMed
Belachew, S., Chittajallu, R., Aguirre, A. A., et al. (2003). Postnatal NG2 proteoglycan-expressing progenitor cells are intrinsically multipotent and generate functional neurons. J Cell Biol 161, 169–186.CrossRefGoogle ScholarPubMed
Ben-Hur, T., Einstein, O., Mizrachi-Kol, R., et al. (2003). Transplanted multipotential neural precursor cells migrate into the inflamed white matter in response to experimental autoimmune encephalomyelitis. Glia 41, 73–80.CrossRefGoogle ScholarPubMed
Bengtsson, S. L., Nagy, Z., Skare, S., Forsman, L., Forssberg, H., and Ullen, F. (2005). Extensive piano practicing has regionally specific effects on white matter development. Nat Neurosci 8, 1148–1150.CrossRefGoogle ScholarPubMed
Benjamins, J. A., Iwata, R., and Hazlett, J. (1978). Kinetics of entry of proteins into the myelin membrane. J Neurochem 31, 1077–1085.CrossRefGoogle ScholarPubMed
Benjamins, J. A., Nedelkoska, L., and George, E. B. (2003). Protection of mature oligodendrocytes by inhibitors of caspases and calpains. Neurochem Res 28, 143–152.CrossRefGoogle ScholarPubMed
Bennett, V. and Lambert, S. (1999). Physiological roles of axonal ankyrins in survival of premyelinated axons and localization of voltage-gated sodium channels. J Neurocytol 28, 303–318.CrossRefGoogle ScholarPubMed
Benninger, Y., Thurnherr, T., Pereira, J. A., et al. (2007). Essential and distinct roles for cdc42 and rac1 in the regulation of Schwann cell biology during peripheral nervous system development. J Cell Biol 177, 1051–1061.CrossRefGoogle ScholarPubMed
Benson, M. D., Romero, M. I., Lush, M. E., Lu, Q. R., Henkemeyer, M., and Parada, L. F. (2005). Ephrin-B3 is a myelin-based inhibitor of neurite outgrowth. Proc Natl Acad Sci USA 102, 10694–10699.CrossRefGoogle ScholarPubMed
Bentley, C. A. and Lee, K. F. (2000). p75 is important for axon growth and Schwann cell migration during development. J Neurosci 20, 7706–7715.CrossRefGoogle ScholarPubMed
Benveniste, H., Drejer, J., Schousboe, A., and Diemer, N. H. (1984). Elevation of the extracellular concentrations of glutamate and aspartate in rat hippocampus during transient cerebral ischemia monitored by intracerebral microdialysis. J Neurochem 43, 1369–1374.CrossRefGoogle ScholarPubMed
Bergamaschi, R., Tonietti, S., Franciotta, D., et al. (2004). Oligoclonal bands in Devic's neuromyelitis optica and multiple sclerosis: differences in repeated cerebrospinal fluid examinations. Mult Scler 10, 2–4.CrossRefGoogle ScholarPubMed
Berger, J. R. (2003).Progressive multifocal leukoencephalopathy in acquired immunodeficiency syndrome: explaining the high incidence and disproportionate frequency of the illness relative to other immunosuppressive conditions. J Neurovirol 9 Suppl 1, 38–41.CrossRefGoogle ScholarPubMed
Berger, P., Niemann, A., and Suter, U. (2006). Schwann cells and the pathogenesis of inherited motor and sensory neuropathies (Charcot-Marie-Tooth disease). Glia 54, 243–257.CrossRefGoogle Scholar
Berger, T., Rubner, P., Schautzer, F., et al. (2003). Antimyelin antibodies as a predictor of clinically definite multiple sclerosis after a first demyelinating event. N Engl J Med 349, 139–145.CrossRefGoogle ScholarPubMed
Berger, T., Walz, W., Schnitzer, J., and Kettenmann, H. (1992). GABA- and glutamate-activated currents in glial cells of the mouse corpus callosum slice. J Neurosci Res 31, 21–27.CrossRefGoogle ScholarPubMed
Berghs, S., Aggujaro, D., Dirkx, R.., et al. (2000). BetaIV spectrin, a new spectrin localized at axon initial segments and nodes of ranvier in the central and peripheral nervous system. J Cell Biol 151, 985–1002.CrossRefGoogle ScholarPubMed
Bergles, D. E., Diamond, J. S., and Jahr, C. E. (1999). Clearance of glutamate inside the synapse and beyond. Curr Opin Neurobiol 9, 293–298.CrossRefGoogle ScholarPubMed
Bergles, D. E., Roberts, J. D., Somogyi, P., and Jahr, C. E. (2000). Glutamatergic synapses on oligodendrocyte precursor cells in the hippocampus. Nature 405, 187–191.CrossRefGoogle ScholarPubMed
Bergoffen, J., Scherer, S. S., Wang, S., et al. (1993). Connexin mutations in X-linked Charcot-Marie-Tooth disease. Science 262, 2039–2042.CrossRefGoogle ScholarPubMed
Bergsteindottir, K., Brennan, A., Jessen, K. R., and Mirsky, R. (1992). In the presence of dexamethasone, gamma interferon induces rat oligodendrocytes to express major histocompatibility complex class II molecules. Proc Natl Acad Sci USA 89, 9054–9058.CrossRefGoogle ScholarPubMed
Bernardo, A., Greco, A., Levi, G., and Minghetti, L. (2003). Differential lipid peroxidation, Mn superoxide, and bcl-2 expression contribute to the maturation-dependent vulnerability of oligodendrocytes to oxidative stress. J Neuropathol Exp Neurol 62, 509–519.CrossRefGoogle ScholarPubMed
Besancon, E., Guo, S., Lok, J., Tymianski, M., and Lo, E. H. (2008). Beyond NMDA and AMPA glutamate receptors: emerging mechanisms for ionic imbalance and cell death in stroke. Trends Pharmacol Sci 29, 268–275.CrossRefGoogle ScholarPubMed
Bethea, J. R. and Dietrich, W. D. (2002). Targeting the host inflammatory response in traumatic spinal cord injury. Curr Opin Neurol 15, 355–360.CrossRefGoogle ScholarPubMed
Bezzi, P., Carmignoto, G., Pasti, L., et al.(1998). Prostaglandins stimulate calcium-dependent glutamate release in astrocytes. Nature 391, 281–285.CrossRefGoogle ScholarPubMed
Bezzi, P., Domercq, M., Brambilla, L., et al. (2001). CXCR4-activated astrocyte glutamate release via TNFalpha: amplification by microglia triggers neurotoxicity. Nat Neurosci 4, 702–710.CrossRefGoogle ScholarPubMed
Bhat, M. A., Rios, J. C., Lu, Y., et al. (2001). Axon–glia interactions and the domain organization of myelinated axons requires neurexin IV/Caspr/Paranodin. Neuron 30, 369–383.CrossRefGoogle ScholarPubMed
Billiards, S. S., Haynes, R. L., Folkerth, R. D., et al. (2008). Myelin abnormalities without oligodendrocyte loss in periventricular leukomalacia. Brain Pathol 18, 153–163.CrossRefGoogle ScholarPubMed
Bitsch, A., Kuhlmann, T., da Costa, C., Bunkowski, S., Polak, T., and Brück, W. (2000a). Tumour necrosis factor alpha mRNA expression in early multiple sclerosis lesions: correlation with demyelinating activity and oligodendrocyte pathology. Glia 29, 366–375.3.0.CO;2-Y>CrossRefGoogle ScholarPubMed
Bitsch, A., Schuchardt, J., Bunkowski, S., Kuhlmann, T., and Bruck, W. (2000b). Acute axonal injury in multiple sclerosis. Correlation with demyelination and inflammation. Brain 123, 1174–1183.CrossRefGoogle ScholarPubMed
Bixby, J. L., Lilien, J., and Reichardt, L. F. (1988). Identification of the major proteins that promote neuronal process outgrowth on Schwann cells in vitro. J Cell Biol 107, 353–361.CrossRefGoogle ScholarPubMed
Bizzozero, O. A., McGarry, J. F., and Lees, M. B. (1986). Acylation of rat brain myelin proteolipid protein with different fatty acids. J Neurochem 47, 772–778.CrossRefGoogle ScholarPubMed
Bjartmar, C. and Trapp, B. D. (2001). Axonal and neuronal degeneration in multiple sclerosis: mechanisms and functional consequences. Curr Opin Neurol 14, 271–278.CrossRefGoogle ScholarPubMed
Blaabjerg, M., Fang, L., Zimmer, J., and Baskys, A. (2003). Neuroprotection against NMDA excitotoxicity by group I metabotropic glutamate receptors is associated with reduction of NMDA stimulated currents. Exp Neurol 183, 573–580.CrossRefGoogle Scholar
Black, J. A., Kocsis, J. D., and Waxman, S. G. (1990). Ion channel organization of the myelinated fiber. Trends Neurosci 13, 48–54.CrossRefGoogle ScholarPubMed
Black, J. A., Waxman, S. G., and Smith, K. J. (2006). Remyelination of dorsal column axons by endogenous Schwann cells restores the normal pattern of Nav1.6 and Kv1.2 at nodes of Ranvier. Brain 129, 1319–1329.CrossRefGoogle ScholarPubMed
Blakemore, W. F. (1969). Schmidt–Lantermann incisures in the central nervous system. J Ultrastruct Res 29, 496–498.CrossRefGoogle ScholarPubMed
Blakemore, W. F. (1974). Pattern of remyelination in the CNS. Nature 249, 577–578.CrossRefGoogle ScholarPubMed
Blakemore, W. F. (1977). Remyelination of CNS axons by Schwann cells transplanted from the sciatic nerve. Nature 266, 68–69.CrossRefGoogle ScholarPubMed
Blakemore, W. F. (1981). Observations on myelination and remyelination in the central nervous system. In Development in the Nervous System, Garrod, D. R. and Feldman, J. D., eds. (Cambridge University Press), pp. 289–308.Google Scholar
Blakemore, W. F. and Keirstead, H. S. (1999). The origin of the remyelinating cells in the central nervous system. J Neuroimmunol 98, 69–76.CrossRefGoogle ScholarPubMed
Blaschuk, K. L., Frost, E. E., and Ffrench-Constant, C. (2000). The regulation of proliferation and differentiation in oligodendrocyte progenitor cells by alphaV integrins. Development 127, 1961–1969.Google ScholarPubMed
Blau, H. M. (2002). A twist of fate. Nature 419, 437.CrossRefGoogle ScholarPubMed
Blight, A. R. (1983). Cellular morphology of chronic spinal cord injury in the cat: analysis of myelinated axons by line-sampling. Neuroscience 10, 521–543.CrossRefGoogle ScholarPubMed
Bo, L., Mork, S., Kong, P. A., Nyland, H., Pardo, C. A., and Trapp, B. D. (1994). Detection of MHC class II-antigens on macrophages and microglia, but not on astrocytes and endothelia in active multiple sclerosis lesions. J Neuroimmunol 51, 135–146.CrossRefGoogle Scholar
Bo, L., Quarles, R. H., Fujita, N., Bartoszewicz, Z., Sato, S., and Trapp, B. D. (1995). Endocytic depletion of L-MAG from CNS myelin in quaking mice. J Cell Biol 131, 1811–1820.CrossRefGoogle ScholarPubMed
Boggs, J. M. (2006). Myelin basic protein: a multifunctional protein. Cell Mol Life Sci 63, 1945–1961.CrossRefGoogle ScholarPubMed
Bogler, O., Wren, D., Barnett, S. C., Land, H., and Noble, M. (1990). Cooperation between two growth factors promotes extended self-renewal and inhibits differentiation of oligodendrocyte-type-2 astrocyte (O-2A) progenitor cells. Proc Natl Acad Sci USA 87, 6368–6372.CrossRefGoogle ScholarPubMed
Boison, D. and Stoffel, W. (1994). Disruption of the compacted myelin sheath of axons of the central nervous system in proteolipid protein-deficient mice. Proc Natl Acad Sci USA 91, 11709–11713.CrossRefGoogle ScholarPubMed
Borges, K., Ohlemeyer, C., Trotter, J., and Kettenmann, H. (1994). AMPA/kainate receptor activation in murine oligodendrocyte precursor cells leads to activation of a cation conductance, calcium influx and blockade of delayed rectifying K+ channels. Neuroscience 63, 135–149.CrossRefGoogle ScholarPubMed
Bosio, A., Bussow, H., Adam, J., and Stoffel, W. (1998). Galactosphingolipids and axono-glial interaction in myelin of the central nervous system. Cell Tissue Res 292, 199–210.CrossRefGoogle ScholarPubMed
Bouvier-Labit, C., Liprandi, A., Monti, G., Pellissier, J. F., and Figarella-Branger, D. (2002). CD44H is expressed by cells of the oligodendrocyte lineage and by oligodendrogliomas in humans. J Neurooncol 60, 127–134.CrossRefGoogle ScholarPubMed
Boyle, M. E., Berglund, E. O., Murai, K. K., Weber, L., Peles, E., and Ranscht, B. (2001). Contactin orchestrates assembly of the septate-like junctions at the paranode in myelinated peripheral nerve. Neuron 30, 385–397.CrossRefGoogle ScholarPubMed
Brady, S. T., Witt, A. S., Kirkpatrick, L. L., et al. (1999). Formation of compact myelin is required for maturation of the axonal cytoskeleton. J Neurosci 19, 7278–7288.CrossRefGoogle ScholarPubMed
Braude, P., Minger, S. L., and Warwick, R. M. (2005). Stem cell therapy: hope or hype? BMJ 330, 1159–1160.CrossRefGoogle ScholarPubMed
Braun, P. E., Sandillon, F., Edwards, A., Matthieu, J.-M., and Privat, A. (1988). Immunocytochemical localization by electron microscopy of 2′,3′-cyclic nucleotide 3′-phosphodiesterase in developing oligodendrocytes of normal and mutant brain. J Neurosci 8, 3057–3066.CrossRefGoogle ScholarPubMed
Bre, M.-H., Kreis, T. E., and Karsenti, E. (1987). Control of microtubule nucleation and stability in Madin-Darby canine kidney cells: the occurrence of noncentrosomal, stable detyrosinated microtubules. J Cell Biol 105, 1283–1296.CrossRefGoogle ScholarPubMed
Bregman, B. S., Kunkel-Bagden, E., Schnell, L., Dai, H. N., Gao, D., and Schwab, M. E. (1995). Recovery from spinal cord injury mediated by antibodies to neurite growth inhibitors. Nature 378, 498–501.CrossRefGoogle ScholarPubMed
Brenner, T., Brocke, S., Szafer, F., et al. (1997). Inhibition of nitric oxide synthase for treatment of experimental autoimmune encephalomyelitis. J Immunol 158, 2940–2946.Google ScholarPubMed
Brierley, C. M., Crang, A. J., Iwashita, Y., et al. (2001). Remyelination of demyelinated CNS axons by transplanted human Schwann cells: the deleterious effect of contaminating fibroblasts. Cell Transplant 10, 305–315.CrossRefGoogle ScholarPubMed
Brinkmann, B. G., Agarwal, A., Sereda, M. W., et al. (2008). Neuregulin-1/ErbB signaling serves distinct functions in myelination of the peripheral and central nervous system. Neuron 59, 581–595.CrossRefGoogle ScholarPubMed
Brophy, P. J. (2001). Axoglial junctions: separate the channels or scramble the message. Curr Biol 11, R555–R557.CrossRefGoogle ScholarPubMed
Brosamle, C., Huber, A. B., Fiedler, M., Skerra, A., and Schwab, M. E. (2000). Regeneration of lesioned corticospinal tract fibers in the adult rat induced by a recombinant, humanized IN-1 antibody fragment. J Neurosci 20, 8061–8068.CrossRefGoogle ScholarPubMed
Brosnan, C. F., Bornstein, M. B., and Bloom, B. R. (1981). The effects of macrophage depletion on the clinical and pathologic expression of experimental allergic encephalomyelitis. J Immunol 126, 614–620.Google ScholarPubMed
Brown, G. C. and Bal-Price, A. (2003). Inflammatory neurodegeneration mediated by nitric oxide, glutamate, and mitochondria. Mol Neurobiol 27, 325–355.CrossRefGoogle Scholar
Brownell, B. and Hughes, J. T. (1962). The distribution of plaques in the cerebrum in multiple sclerosis. J Neurol Neurosurg Psychiatry 25, 315–320.CrossRefGoogle ScholarPubMed
Brück, W., Schmied, M., Suchanek, G., et al. (1994). Oligodendrocytes in the early course of multiple sclerosis. Ann Neurol 35, 65–73.CrossRefGoogle ScholarPubMed
Brunner, C., Lassmann, H., Waehneldt, T. V., Matthieu, J. M., and Linington, C. (1989). Differential ultrastructural localization of myelin basic protein, myelin/oligodendroglial glycoprotein, and 2′,3′-cyclic nucleotide 3′-phosphodiesterase in the CNS of adult rats. J Neurochem 52, 296–304.CrossRefGoogle ScholarPubMed
Brushart, T. M. E. (1990). Preferential motor reinnervation: a sequential double-labeling study. Restor Neurol Neurosci281–287.
Brushart, T. M. E. (1993). Motor axons preferentially reinnervate motor pathways. J Neurosci 13, 2730–2738.CrossRefGoogle ScholarPubMed
Brustle, O., Jones, K. N., Learish, R. D., et al. (1999). Embryonic stem cell-derived glial precursors: a source of myelinating transplants. Science 285, 754–756.Google ScholarPubMed
Bruzzone, R., White, T. W., Scherer, S. S., Fischbeck, K. H., and Paul, D. L. (1994). Null mutations of connexin32 in patients with X-linked Charcot-Marie-Tooth disease. Neuron 13, 1253–1260.CrossRefGoogle ScholarPubMed
Bsibsi, M., Ravid, R., Gveric, D., and Noort, J. M. (2002). Broad expression of Toll-like receptors in the human central nervous system. J Neuropathol Exp Neurol 61, 1013–1021.CrossRefGoogle ScholarPubMed
Buffo, A., Zagrebelsky, M., Huber, A. B., et al. (2000). Application of neutralizing antibodies against NI-35/250 myelin-associated neurite growth inhibitory proteins to the adult rat cerebellum induces sprouting of uninjured Purkinje cell axons. J Neurosci 20, 2275–2286.CrossRefGoogle ScholarPubMed
Bunge, M. B. and Pearse, D. D. (2003). Transplantation strategies to promote repair of the injured spinal cord. J Rehabil Res Dev 40, 55–62.CrossRefGoogle ScholarPubMed
Bunge, M. B.Bunge, R. P., and Ris, H. (1961). Ultrastructural study of remyelination in an experimental lesion in adult cat spinal cord. J Biophys Biochem Cytol 10, 67–94.CrossRefGoogle Scholar
Bunge, R. P., Bunge, M. B., and Ris, H. (1960). Electron microscopic study of demyelination in an experimentally induced lesion in adult cat spinal cord. J Biophys Biochem Cytol 7, 685–696.CrossRefGoogle Scholar
Buntinx, M., Moreels, M., Vandenabeele, F., et al. (2004). Cytokine-induced cell death in human oligodendroglial cell lines: I. Synergistic effects of IFN-gamma and TNF-alpha on apoptosis. J Neurosci Res 76, 834–845.CrossRefGoogle ScholarPubMed
Burgmaier, G., Schönrock, M. L., Kuhlmann, T., Richter-Landsberg, C., and Brück, W. (2000). Association of increased bcl-2 expression with rescue from TNF-α induced cell death in the oligodendrocyte cell line OLN-93. J Neurochem 75, 2270–2276.CrossRefGoogle ScholarPubMed
Burke, R. E. (2007). Sir Charles Sherrington's the integrative action of the nervous system: a centenary appreciation. Brain 130, 887–894.CrossRefGoogle ScholarPubMed
Burnashev, N., Monyer, H., Seeburg, P. H., and Sakmann, B. (1992). Divalent ion permeability of AMPA receptor channels is dominated by the edited form of a single subunit. Neuron 8, 189–198.CrossRefGoogle ScholarPubMed
Burns, J., Rosenzweig, A., Zweiman, B., and Lisak, R. (1983). Isolation of myelin basic protein-reactive T-cell lines from normal human blood. Cell Immunol 81, 435–440.CrossRefGoogle ScholarPubMed
Burt, R. K., Loh, Y., Pearce, W., et al. (2008). Clinical applications of blood-derived and marrow-derived stem cells for nonmalignant diseases. J Am Med Assoc 299, 925–936.CrossRefGoogle ScholarPubMed
Bushati, N. and Cohen, S. M. (2007). microRNA functions. Annu Rev Cell Dev Biol 23, 175–205.CrossRefGoogle ScholarPubMed
Butt, A. M., Duncan, A., Hornby, M. F., et al. (1999). Cells expressing the NG2 antigen contact nodes of Ranvier in adult CNS white matter. Glia 26, 84–91.3.0.CO;2-L>CrossRefGoogle ScholarPubMed
Butt, A. M., Hamilton, N., Hubbard, P., Pugh, M., and Ibrahim, M. (2005). Synantocytes: the fifth element. J Anat 207, 695–706.CrossRefGoogle ScholarPubMed
Butt, A. M., Kiff, J., Hubbard, P., and Berry, M. (2002). Synantocytes: new functions for novel NG2 expressing glia. J Neurocytol 31, 551–565.CrossRefGoogle ScholarPubMed
Butzkueven, H., Emery, B., Cipriani, T., Marriott, M. P., and Kilpatrick, T. J. (2006). Endogenous leukemia inhibitory factor production limits autoimmune demyelination and oligodendrocyte loss. Glia 53, 696–703.CrossRefGoogle ScholarPubMed
Byravan, S., Foster, L. M., Phan, T., Verity, A. N., and Campagnoni, A. T. (1994). Murine oligodendroglial cells express nerve growth factor. Proc Natl Acad Sci USA 91, 8812–8816.CrossRefGoogle ScholarPubMed
Cai, D., Qiu, J., Cao, Z., McAtee, M., Bregman, B. S., and Filbin, M. T. (2001). Neuronal cyclic AMP controls the developmental loss in ability of axons to regenerate. J Neurosci 21, 4731–4739.CrossRefGoogle ScholarPubMed
Cai, J., Qi, Y., Hu, X., et al. (2005). Generation of oligodendrocyte precursor cells from mouse dorsal spinal cord independent of Nkx6 regulation and Shh signaling. Neuron 45, 41–53.CrossRefGoogle ScholarPubMed
Cai, Z., Pan, Z. L., Pang, Y., Evans, O. B., and Rhodes, P. G. (2000). Cytokine induction in fetal rat brains and brain injury in neonatal rats after maternal lipopolysaccharide administration. Pediatr Res 47, 64–72.CrossRefGoogle ScholarPubMed
Caldwell, J. H., Schaller, K. L., Lasher, R. S., Peles, E., and Levinson, S. R. (2000). Sodium channel Na(v)1.6 is localized at nodes of Ranvier, dendrites, and synapses. Proc Natl Acad Sci USA 97, 5616–5620.CrossRefGoogle ScholarPubMed
Cameron-Curry, P. and Douarin, N. M. (1995). Oligodendrocyte precursors originate from both the dorsal and the ventral parts of the spinal cord. Neuron 15, 1299–1310.CrossRefGoogle ScholarPubMed
Campagnoni, A. T., Pribyl, T. M., Campagnoni, C. W., et al. (1993). Structure and developmental regulation of Golli-mbp, a 105 kilobase gene that encompasses the myelin basic protein gene and is expressed in cells in the oligodendrocytes lineage in the brain. J Biol Chem 268, 4930–4938.Google Scholar
Cannella, B. and Raine, C. S. (2004). Multiple sclerosis: cytokine receptors on oligodendrocytes predict innate regulation. Ann. Neurol. 55, 46–57.CrossRefGoogle ScholarPubMed
Cannella, B., Gaupp, S., Omari, K. M., and Raine, C. S. (2007). Multiple sclerosis: death receptor expression and oligodendrocyte apoptosis in established lesions. J Neuroimmunol 188, 128–137.CrossRefGoogle ScholarPubMed
Cao, Q., Zhang, Y. P., Iannotti, C., et al. (2005). Functional and electrophysiological changes after graded traumatic spinal cord injury in adult rat. Exp Neurol 191 Suppl 1, S3–S16.CrossRefGoogle ScholarPubMed
Carbonetto, S., Evans, D., and Cochard, P. (1987). Nerve fiber growth in culture on tissue substrata from central and peripheral nervous systems. J Neurosci 7, 610–620.CrossRefGoogle ScholarPubMed
Carenini, S., Montag, D., Cremer, H., Schachner, M., and Martini, R. (1997). Absence of the myelin-associated glycoprotein (MAG) and the neural cell adhesion molecule (N-CAM) interferes with the maintenance, but not with the formation of peripheral myelin. Cell Tissue Res 287, 3–9.CrossRefGoogle Scholar
Carnegie, P. R., Dunkley, P. R., Kemp, B. E., and Murray, A. W. (1974). Phosphorylation of selected serine and threonine residues in myelin basic protein by endogenous and exogenous protein kinases. Nature 249, 147–150.CrossRefGoogle ScholarPubMed
Caroni, P. and Schwab, M. E. (1988a). Antibody against myelin-associated inhibitor of neurite growth neutralizes nonpermissive substrate properties of CNS white matter. Neuron 1, 85–96.CrossRefGoogle ScholarPubMed
Caroni, P. and Schwab, M. E. (1988b). Two membrane protein fractions from rat central myelin with inhibitory properties for neurite growth and fibroblast spreading. J Cell Biol 106, 1281–1288.CrossRefGoogle ScholarPubMed
Carroll, S. L., Miller, M. L., Frohnert, P. W., Kim, S. S., and Corbett, J. A. (1997). Expression of neuregulins and their putative receptors, ErbB2 and ErbB3, is induced during Wallerian degeneration. J Neurosci 17, 1642–1659.CrossRefGoogle ScholarPubMed
Carroll, W. M. and Jennings, A. R. (1994). Early recruitment of oligodendrocyte precursors in CNS demyelination. Brain 117, 563–578.CrossRefGoogle ScholarPubMed
Carroll, W. M., Jennings, A. R., and Ironside, L. J. (1998). Identification of the adult resting progenitor cell by autoradiographic tracking of oligodendrocyte precursors in experimental CNS demyelination. Brain 121 (Pt 2), 293–302.CrossRefGoogle ScholarPubMed
Carson, J. H., Cui, H., and Barbarese, E. (2001). The balance of power in RNA trafficking. Curr Opin Neurobiol 11, 558–563.CrossRefGoogle ScholarPubMed
Carson, J. H., Gao, Y., Tatavarty, V., et al. (2008). Multiplexed RNA trafficking in oligodendrocytes and neurons. Biochim Biophys Acta 1779, 453–458.CrossRefGoogle ScholarPubMed
Carson, J. H., Worboys, K., Ainger, K., and Barbarese, E. (1997). Translocation of myelin basic protein mRNA in oligodendrocytes requires microtubules and kinesin. Cell Motil Cytoskeleton 38, 318–328.3.0.CO;2-#>CrossRefGoogle ScholarPubMed
Carson, M. J., Behringer, R. R., Brinster, R. L., and McMorris, F. A. (1993). Insulin-like growth factor I increases brain growth and central nervous system myelination in transgenic mice. Neuron 10, 729–740.CrossRefGoogle ScholarPubMed
Casha, S., Yu, W. R., and Fehlings, M. G. (2001). Oligodendroglial apoptosis occurs along degenerating axons and is associated with FAS and p75 expression following spinal cord injury in the rat. Neuroscience 103, 203–218.CrossRefGoogle ScholarPubMed
Cassiani-Ingoni, R., Greenstone, H. L., Donati, D., et al. (2005). CD46 on glial cells can function as a receptor for viral glycoprotein-mediated cell-cell fusion. Glia 52, 252–258.CrossRefGoogle ScholarPubMed
Chan, J. A., Krichevsky, A. M., and Kosik, K. S. (2005). MicroRNA-21 is an antiapoptotic factor in human glioblastoma cells. Cancer Res 65, 6029–6033.CrossRefGoogle ScholarPubMed
Chan, J. R., Watkins, T. A., Cosgaya, J. M., et al. (2004). NGF controls axonal receptivity to myelination by Schwann cells or oligodendrocytes. Neuron 43, 183–191.CrossRefGoogle ScholarPubMed
Chang, A., Nishiyama, A., Peterson, J., Prineas, J., and Trapp, B. D. (2000). NG2-positive oligodendrocyte progenitor cells in adult human brain and multiple sclerosis lesions. J Neurosci 20, 6404–6412.CrossRefGoogle ScholarPubMed
Chang, A., Tourtellotte, W. W., Rudick, R., and Trapp, B. D. (2002). Premyelinating oligodendrocytes in chronic lesions of multiple sclerosis. New Engl J Med 346, 165–173.CrossRefGoogle ScholarPubMed
Charcot, J. (1868). Histologie de la sclérose en plaque. Gaz Hop Civ Mil Empire Ottoman 41, 554–566.Google Scholar
Charcot, J M. (1877). Lecture VI. Disseminated sclerosis. Pathological anatomy. In Lectures on The Diseases of the Nervous System, Trans. Sigerson, George. (London: The New Sydenham Society), pp. 157–181.Google Scholar
Charles, P., Hernandez, M. P., Stankoff, B., et al. (2000). Negative regulation of central nervous system myelination by polysialylated-neural cell adhesion molecule. Proc Natl Acad Sci USA 97, 7585–7590.CrossRefGoogle ScholarPubMed
Charles, P., Reynolds, R., Seilhean, D., et al. (2002a). Re-expression of PSA-NCAM by demyelinated axons: an inhibitor of remyelination in multiple sclerosis? Brain 125, 1972–1979.CrossRefGoogle ScholarPubMed
Charles, P., Tait, S., Faivre-Sarrailh, C., et al. (2002b). Neurofascin is a glial receptor for the paranodin/Caspr-contactin axonal complex at the axoglial junction. Curr Biol 12, 217–220.CrossRefGoogle ScholarPubMed
Charo, I. F. and Ransohoff, R. M. (2006). The many roles of chemokines and chemokine receptors in inflammation. N Engl J Med 354, 610–621.CrossRefGoogle ScholarPubMed
Chaudhry, N. and Filbin, M. T. (2007). Myelin-associated inhibitory signaling and strategies to overcome inhibition. J Cereb Blood Flow Metab 27, 1096–1107.CrossRefGoogle ScholarPubMed
Chaudhuri, A. (2006). Lessons for clinical trials from natalizumab in multiple sclerosis. BMJ 332, 416–419.CrossRefGoogle ScholarPubMed
Chaudhuri, A. and Behan, P. O. (2005). Multiple sclerosis: looking beyond autoimmunity. J R Soc Med 98, 303–306.CrossRefGoogle ScholarPubMed
Chen, J. T., Kuhlmann, T., Jansen, G. H., et al. (2007a). Voxel-based analysis of the evolution of magnetization transfer ratio to quantify remyelination and demyelination with histopathological validation in a multiple sclerosis lesion. Neuroimage 36, 1152–1158.CrossRefGoogle Scholar
Chen, K. and Rajewsky, N. (2006). Natural selection on human microRNA binding sites inferred from SNP data. Nat Genet 38, 1452–1456.CrossRefGoogle ScholarPubMed
Chen, M. S., Huber, A. B., Haar, M. E., et al. (2000). Nogo-A is a myelin-associated neurite outgrowth inhibitor and an antigen for monoclonal antibody IN-1. Nature 403, 434–439.CrossRefGoogle ScholarPubMed
Chen, Q., Long, Y., Yuan, X., et al. (2005). Protective effects of bone marrow stromal cell transplantation in injured rodent brain: synthesis of neurotrophic factors. J Neurosci Res 80, 611–619.CrossRefGoogle ScholarPubMed
Chen, W., Mahadomrongkul, V., Berger, U. V., et al. (2004). The glutamate transporter GLT1a is expressed in excitatory axon terminals of mature hippocampal neurons. J Neurosci 24, 1136–1148.CrossRefGoogle ScholarPubMed
Chen, Z. L., Yu, W. M., and Strickland, S. (2007b). Peripheral regeneration. Annu Rev Neurosci 30, 209–233.CrossRefGoogle ScholarPubMed
Cheong, K. H., Zacchetti, D., Schneeberger, E. E., and Simons, K. (1999). VIP17/MAL, a lipid raft-associated protein, is involved in apical transport in MDCK cells. Proc Natl Acad Sci USA 96, 6241–6248.CrossRefGoogle ScholarPubMed
Chew, L. J., King, W. C., Kennedy, A., and Gallo, V. (2005). Interferon-gamma inhibits cell cycle exit in differentiating oligodendrocyte progenitor cells. Glia 52, 127–143.CrossRefGoogle ScholarPubMed
Chitnis, T., Imitola, J., Wang, Y., et al. (2007). Elevated neuronal expression of CD200 protects Wlds mice from inflammation-mediated neurodegeneration. Am J Pathol 170, 1695–1712.CrossRefGoogle ScholarPubMed
Chivatakarn, O., Kaneko, S., He, Z., Tessier-Lavigne, M., and Giger, R. J. (2007). The Nogo-66 receptor NgR1 is required only for the acute growth cone-collapsing but not the chronic growth-inhibitory actions of myelin inhibitors. J Neurosci 27, 7117–7124.CrossRefGoogle Scholar
Choi, D. W. (1988). Glutamate neurotoxicity and diseases of the nervous system. Neuron 1, 623–634.CrossRefGoogle ScholarPubMed
Choi, J., Opalenik, S. R., Wu, W. C., Thompson, J. A., and Forman, H. J. (2000). Modulation of glutathione synthetic enzymes by acidic fibroblast growth factor. Arch Biochem Biophys 375, 201–209.CrossRefGoogle ScholarPubMed
Chopp, M., Zhang, X. H., Li, Y., et al. (2000). Spinal cord injury in rat: treatment with bone marrow stromal cell transplantation. NeuroReport 11, 3001–3005.CrossRefGoogle ScholarPubMed
Chow, E., Mottahedeh, J., Prins, M., Ridder, W., Nusinowitz, S., and Bronstein, J. M. (2005). Disrupted compaction of CNS myelin in an OSP/Claudin-11 and PLP/DM20 double knockout mouse. Mol Cell Neurosci 29, 405–413.CrossRefGoogle Scholar
Chun, S. J., Rasband, M. N., Sidman, R. L., Habib, A. A., and Vartanian, T. (2003). Integrin-linked kinase is required for laminin-2-induced oligodendrocyte cell spreading and CNS myelination. J Cell Biol 163, 397–408.CrossRefGoogle ScholarPubMed
Ciutat, D., Caldero, J., Oppenheim, R. W., and Esquerda, J. E. (1996). Schwann cell apoptosis during normal development and after axonal degeneration induced by neurotoxins in the chick embryo. J Neurosci 16, 3979–3990.CrossRefGoogle ScholarPubMed
Cleveland, D. W. (1987). The multitubulin hypothesis revisited: what have we learned? J. Cell Biol 104, 381–383.CrossRefGoogle Scholar
Cogle, C. R., Yachnis, A. T., Laywell, E. D., et al. (2004). Bone marrow transdifferentiation in brain after transplantation: a retrospective study. Lancet 363, 1432–1437.CrossRefGoogle ScholarPubMed
Cohen, N. R., Taylor, J. S., Scott, L. B., Guillery, R. W., Soriano, P., and Furley, A. J. (1998). Errors in corticospinal axon guidance in mice lacking the neural cell adhesion molecule L1. Curr Biol 8, 26–33.CrossRefGoogle ScholarPubMed
Colello, R. J., Pott, U., and Schwab, M. E. (1994). The role of oligodendrocytes and myelin on axon maturation in the developing rat retinofugal pathway. J Neurosci 14, 2594–2605.CrossRefGoogle ScholarPubMed
Coleman, M. P. and Perry, V. H. (2002). Axon pathology in neurological disease: a neglected therapeutic target. Trends Neurosci 25, 532–537.CrossRefGoogle ScholarPubMed
Coles, A., Wing, M. G., Molyneux, P., Paolillo, A., Davie, C. A., and Hale, G. (1999). Monoclonal antibody treatment exposes three mechanisms underlying the clinical course of multiple sclerosis. Ann Neurol 46, 304.3.0.CO;2-#>CrossRefGoogle ScholarPubMed
Colman, D. R., Kreibich, G., Frey, A. B., and Sabatini, D. D. (1982). Synthesis and incorporation of myelin polypeptide into CNS myelin. J Cell Biol 95, 598–608.CrossRefGoogle Scholar
Coman, I., Aigrot, M. S., Seilhean, D., et al. (2006). Nodal, paranodal and juxtaparanodal axonal proteins during demyelination and remyelination in multiple sclerosis. Brain 129, 3186–3195.CrossRefGoogle ScholarPubMed
Coman, I., Barbin, G., Charles, P., Zalc, B., and Lubetzki, C. (2005). Axonal signals in central nervous system myelination, demyelination and remyelination. J Neurol Sci 233, 67–71.CrossRefGoogle ScholarPubMed
Compston, D. A. S. (1996). Remyelination of the central nervous system. Mult Scler 1, 388–392.CrossRefGoogle ScholarPubMed
Compston, D. A. S., Morgan, B. P., Campbell, A. K., et al. (1989). Immunocytochemical localization of the terminal complement complex in multiple sclerosis. Neuropathol Appl Neurobiol 15, 307–316.CrossRefGoogle ScholarPubMed
Conde, J. R. and Streit, W. J. (2006). Microglia in the aging brain. J Neuropathol Exp Neurol 65, 199–203.CrossRefGoogle ScholarPubMed
Confavreux, C., Vukusic, S., Moreau, T., and Adeleine, P. (2000). Relapses and progression of disability in multiple sclerosis. N Engl J Med 343, 1430–1438.CrossRefGoogle ScholarPubMed
Conn, J. P. and Patel, J. (1994). The Metabotropic Glutamate Receptors. (Totowa, NJ: Humana Press).CrossRefGoogle Scholar
Connor, J. R. and Menzies, S. L. (1996). Relationship of iron to oligodendrocytes and myelination. Glia 17, 83–93.3.0.CO;2-7>CrossRefGoogle ScholarPubMed
Corfas, G., Velardez, M. O., Ko, C. P., Ratner, N., and Peles, E. (2004). Mechanisms and roles of axon-Schwann cell interactions. J Neurosci 24, 9250–9260.CrossRefGoogle ScholarPubMed
Cornbrooks, C. J., Carey, D. J., McDonald, J. A., Timpl, R., and Bunge, R. P. (1983). In vivo and in vitro observations on laminin production by Schwann cells. Proc Natl Acad Sci USA 80, 3850–3854.CrossRefGoogle ScholarPubMed
Counsell, S. J., Allsop, J. M., Harrison, M. C., et al. (2003). Diffusion-weighted imaging of the brain in preterm infants with focal and diffuse white matter abnormality. Pediatrics 112, 1–7.CrossRefGoogle ScholarPubMed
Craig, A., Ling Luo, N., Beardsley, D. J., et al. (2003). Quantitative analysis of perinatal rodent oligodendrocyte lineage progression and its correlation with human. Exp Neurol 181, 231–240.CrossRefGoogle ScholarPubMed
Craner, M. J., Lo, A. C., Black, J. A., and Waxman, S. G. (2003). Abnormal sodium channel distribution in optic nerve axons in a model of inflammatory demyelination. Brain 126, 1552–1561.CrossRefGoogle Scholar
Craner, M. J., Newcombe, J., Black, J. A., Hartle, C., Cuzner, M. L., and Waxman, S. G. (2004). Molecular changes in neurons in multiple sclerosis: altered axonal expression of Nav1.2 and Nav1.6 sodium channels and Na+/Ca2+ exchanger. Proc Natl Acad Sci USA 101, 8168–8173.CrossRefGoogle Scholar
Crespo, D., Asher, R. A., Lin, R., Rhodes, K. E., and Fawcett, J. W. (2007). How does chondroitinase promote functional recovery in the damaged CNS? Exp Neurol 206, 159–171.CrossRefGoogle ScholarPubMed
Cross, A. H., Manning, P. T., Stern, M. K., and Misko, T. P. (1997). Evidence for the production of peroxynitrite in inflammatory CNS demyelination. J Neuroimmunol 80, 121–130.CrossRefGoogle ScholarPubMed
Cross, A. H., Trotter, J. L., and Lyons, J.-A. (2001). B cells and antibodies in CNS demyelinating disease. J Neuroimmunol 112, 1–14.CrossRefGoogle ScholarPubMed
Cross, D., Farias, G., Dominguez, J., Avila, J., and Maccioni, R. B. (1994). Carboxyl terminal sequences of beta-tubulin involved in the interaction of HMW-MAPs. Studies using site-specific antibodies. Mol Cell Biochem 132, 81–90.CrossRefGoogle ScholarPubMed
Crow, J. P., Ye, Y. Z., Strong, M., Kirk, M., Barnes, S., and Beckman, J. S. (1997). Superoxide dismutase catalyzes nitration of tyrosines by peroxynitrite in the rod and head domains of neurofilament-L. J Neurochem 69, 1945–1953.CrossRefGoogle ScholarPubMed
Crowe, M. J., Bresnahan, J. C., Shuman, S. L., Masters, J. N., and Beattie, M. S. (1997). Apoptosis and delayed degeneration after spinal cord injury in rats and monkeys. Nat Med 3, 73–76.CrossRefGoogle ScholarPubMed
Cudrici, C., Niculescu, F., Jensen, T., et al. (2006). C5b-9 terminal complex protects oligodendrocytes from apoptotic cell death by inhibiting caspase-8 processing and up-regulating FLIP. J Immunol 176, 3173–3180.CrossRefGoogle ScholarPubMed
Cummings, B. J., Uchida, N., Tamaki, S. J., et al. (2005). Human neural stem cells differentiate and promote locomotor recovery in spinal cord-injured mice. Proc Natl Acad Sci USA 102, 14069–14074.CrossRefGoogle ScholarPubMed
Cuzner, M. L., Loughlin, A. J., Mosley, K., and Woodroofe, M. N. (1994). The role of microglia macrophages in the processes of inflammatory demyelination and remyelination. Neuropathol Appl Neurobiol 20, 200–201.Google ScholarPubMed
D'Antoni, S., Berretta, A., Bonaccorso, C. M., et al. (2008). Metabotropic glutamate receptors in glial cells. Neurochem Res 33, 2436–2443.CrossRefGoogle ScholarPubMed
da Cunha, A., Jefferson, J. A., Jackson, R. W., and Vitkovic, L. (1993). Glial cell-specific mechanisms of TGF-beta 1 induction by IL-1 in cerebral cortex. J. Neuroimmunol. 42, 71–85.CrossRefGoogle ScholarPubMed
Dahme, M., Bartsch, U., Martini, R., Anliker, B., Schachner, M., and Mantei, N. (1997). Disruption of the mouse L1 gene leads to malformations of the nervous system. Nat Genet 17, 346–349.CrossRefGoogle Scholar
Dai, X., Lercher, L. D., Clinton, P. M., et al. (2003). The trophic role of oligodendrocytes in the basal forebrain. J Neurosci 23, 5846–5853.CrossRefGoogle ScholarPubMed
Dalitz, P., Harding, R., Rees, S. M., and Cock, M. L. (2003). Prolonged reductions in placental blood flow and cerebral oxygen delivery in preterm fetal sheep exposed to endotoxin: possible factors in white matter injury after acute infection. J Soc Gynecol Investig 10, 283–290.Google ScholarPubMed
Dammann, O. and Leviton, A. (2000). Role of the fetus in perinatal infection and neonatal brain damage. Curr Opin Pediatr 12, 99–104.CrossRefGoogle ScholarPubMed
Dammann, O., Kuban, K. C., and Leviton, A. (2002). Perinatal infection, fetal inflammatory response, white matter damage, and cognitive limitations in children born preterm. Men Retard Dev Dis 8, 46–50.CrossRefGoogle ScholarPubMed
Danbolt, N. C. (2001). Glutamate uptake. Prog Neurobiol 65, 1–105.CrossRefGoogle ScholarPubMed
Dashiell, S. M., Tanner, S. L., Pant, H. C., and Quarles, R. H. (2002). Myelin-associated glycoprotein modulates expression and phosphorylation of neuronal cytoskeletal elements and their associated kinases. J Neurochem 81, 1263–1272.CrossRefGoogle ScholarPubMed
David, S. and Aguayo, A. J. (1981). Axonal elongation into peripheral nervous system “bridges” after central nervous system injury in adult rats. Science 214, 931–933.CrossRefGoogle ScholarPubMed
David, S., Braun, P. E., Jackson, D. L., Kottis, V., and McKerracher, L. (1995). Laminin overrides the inhibitory effects of peripheral nervous system and central nervous system myelin-derived inhibitors of neurite growth. J Neurosci Res 42, 594–602.CrossRefGoogle ScholarPubMed
Davis, A. D., Weatherby, T. M., Hartline, D. K., and Lenz, P. H. (1999). Myelin-like sheaths in copepod axons. Nature 398, 571.CrossRefGoogle ScholarPubMed
Davis, J. Q., Lambert, S., and Bennett, V. (1996). Molecular composition of the node of Ranvier: identification of ankyrin-binding cell adhesion molecules neurofascin (mucin+/third FNIII domain–) and NrCAM at nodal axon segments. J Cell Biol 135(5), 1355–1367.CrossRefGoogle ScholarPubMed
Dawson, J. W. (1916). The histology of disseminated sclerosis. Edinb Med J 17, 229–410.Google Scholar
Dawson, M. R., Levine, J. M., and Reynolds, R. (2000). NG2-expressing cells in the central nervous system: are they oligodendroglial progenitors?J Neurosci Res 61, 471–479.3.0.CO;2-N>CrossRefGoogle ScholarPubMed
Dawson, M. R., Polito, A., Levine, J. M., and Reynolds, R. (2003). NG2-expressing glial progenitor cells: an abundant and widespread population of cycling cells in the adult rat CNS. Mol Cell Neurosci 24, 476–488.CrossRefGoogle ScholarPubMed
Vries, H. and Hoekstra, D. (2000). On the biogenesis of the myelin sheath: cognate polarized trafficking pathways in oligodendrocytes. Glycoconj J 17, 181–190.CrossRefGoogle ScholarPubMed
Waegh, S. M., Lee, V. M., and Brady, S. T. (1992). Local modulation of neurofilament phosphorylation, axonal caliber, and slow axonal transport by myelinating Schwann cells. Cell 68, 451–463.CrossRefGoogle ScholarPubMed
Decker, L., Avellana-Adalid, V., Nait-Oumesmar, B., Durbec, P., and Baron-Van Evercooren, A. (2000). Oligodendrocyte precursor migration and differentiation: combined effects of PSA residues, growth factors, and substrates. Mol Cell Neurosci 16, 422–439.CrossRefGoogle ScholarPubMed
Delarasse, C., Daubas, P., Mars, L. T., et al. (2003). Myelin/oligodendrocyte glycoprotein-deficient (MOG-deficient) mice reveal lack of immune tolerance to MOG in wild-type mice. J Clin Invest 112, 544–553.CrossRefGoogle ScholarPubMed
Deloire-Grassin, M. S., Brochet, B., Quesson, B., et al. (2000). In vivo evaluation of remyelination in rat brain by magnetization transfer imaging. J Neurol Sci 178, 10–16.CrossRefGoogle ScholarPubMed
Demerens, C., Stankoff, B., Logak, M., et al. (1996). Induction of myelination in the central nervous system by electrical activity. Proc Natl Acad Sci USA 93, 9887–9892.CrossRefGoogle ScholarPubMed
Deng, W., Rosenberg, P. A., Volpe, J. J., and Jensen, F. E. (2003). Calcium-permeable AMPA/kainate receptors mediate toxicity and preconditioning by oxygen-glucose deprivation in oligodendrocyte precursors. Proc Natl Acad Sci USA 100, 6801–6806.CrossRefGoogle ScholarPubMed
Deng, W., Wang, H., Rosenberg, P. A., Volpe, J. J., and Jensen, F. E. (2004). Role of metabotropic glutamate receptors in oligodendrocyte excitotoxicity and oxidative stress. Proc Natl Acad Sci USA 101, 7751–7756.CrossRefGoogle ScholarPubMed
Denisenko-Nehrbass, N., Oguievetskaia, K., Goutebroze, L., et al. (2003). Protein 4.1B associates with both Caspr/paranodin and Caspr2 at paranodes and juxtaparanodes of myelinated fibres. Eur J Neurosci 17, 411–416.CrossRefGoogle ScholarPubMed
Dermietzel, R., Traub, O., Hwang, T. K., et al. (1989). Differential expression of three gap junction proteins in developing and mature brain tissues. Proc Natl Acad Sci USA 86, 10148–10152.CrossRefGoogle ScholarPubMed
DeSilva, T. M., Kabakov, A. Y., Goldhoff, P. E., Volpe, J. J., and Rosenberg, P. A. (2009). Regulation of glutamate transport in developing rat oligodendrocytes. J Neurosci 29, 7898–7908.CrossRefGoogle ScholarPubMed
DeSilva, T. M., Kinney, H. C., Borenstein, N. S., et al. (2007). The glutamate transporter EAAT2 is transiently expressed in developing human cerebral white matter. J Comp Neurol 501, 879–890.CrossRefGoogle ScholarPubMed
Devaux, J. J. and Scherer, S. S. (2005). Altered ion channels in an animal model of Charcot-Marie-Tooth disease type IA. J Neurosci 25, 1470–1480.CrossRefGoogle Scholar
Dhaunchak, A. S. and Nave, K. A. (2007). A common mechanism of PLP/DM20 misfolding causes cysteine-mediated endoplasmic reticulum retention in oligodendrocytes and Pelizaeus-Merzbacher disease. Proc Natl Acad Sci USA 104, 17813–17818.CrossRefGoogle ScholarPubMed
Diers-Fenger, M., Kirchhoff, F., Kettenmann, H., Levine, J. M., and Trotter, J. (2001). AN2/NG2 protein-expressing glial progenitor cells in the murine CNS: isolation, differentiation, and association with radial glia. Glia 34, 213–228.CrossRefGoogle ScholarPubMed
Dimos, J. T., Rodolfa, K. T., Niakan, K. K., et al. (2008). Induced pluripotent stem cells generated from patients with ALS can be differentiated into motor neurons. Science 321, 1218–1221.CrossRefGoogle ScholarPubMed
Dimou, L., Schnell, L., Montani, L., et al. (2006). Nogo-A-deficient mice reveal strain-dependent differences in axonal regeneration. J Neurosci 26, 5591–5603.CrossRefGoogle ScholarPubMed
Dingledine, R., Borges, K., Bowie, D., and Traynelis, S. F. (1999). The glutamate receptor ion channels. Pharmacol Rev 51, 7–61.Google ScholarPubMed
Disanza, A., Steffen, A., Hertzog, M., Frittoli, E., Rottner, K., and Scita, G. (2005). Actin polymerization machinery: the finish line of signaling networks, the starting point of cellular movement. Cell Mol Life Sci 62, 955–970.CrossRefGoogle ScholarPubMed
Dobbing, J. and Sands, J. (1979). Comparative aspects of the brain growth spurt. Early Hum Dev 3, 79–83.CrossRefGoogle ScholarPubMed
Dolei, A. and Perron, H. (2009). The multiple sclerosis-associated retrovirus and its HERV-W endogenous family: a biological interface between virology, genetics, and immunology in human physiology and disease. J Neurovirol 15, 4–13.CrossRefGoogle ScholarPubMed
Domeniconi, M., Cao, Z., Spencer, T., et al. (2002). Myelin-associated glycoprotein interacts with the Nogo66 receptor to inhibit neurite outgrowth. Neuron 35, 283–290.CrossRefGoogle ScholarPubMed
Domercq, M., Sanchez-Gomez, M. V., Sherwin, C., Etxebarria, E., Fern, R., and Matute, C. (2007). System xc- and glutamate transporter inhibition mediates microglial toxicity to oligodendrocytes. J Immunol 178, 6549–6556.CrossRefGoogle ScholarPubMed
Dong, Z., Brennan, A., Liu, N., et al. (1995). Neu differentiation factor is a neuron-glia signal and regulates survival, proliferation, and maturation of rat Schwann cell precursors. Neuron 15, 585–596.CrossRefGoogle ScholarPubMed
Dorr, J., Bechmann, I., Waiczies, S., et al. (2002). Lack of tumor necrosis factor-related apoptosis-inducing ligand but presence of its receptors in the human brain. J Neurosci 22, RC209–RC211.CrossRefGoogle ScholarPubMed
Dowling, P., Husar, W., Menonna, J., Donnenfeld, H., Cook, S., and Sidhu, M. (1997). Cell death and birth in multiple sclerosis brain. J Neurol Sci 149, 1–11.CrossRefGoogle ScholarPubMed
D'Souza, S. D., Bonetti, B., Balasingam, V., Cashman, N. R., Barker, P. A., Troutt, A. B., Raine, C. S., and Antel, J P.et al. (1996). Multiple sclerosis: Fas signaling in oligodendrocyte cell death. J Exp Med 184, 2361–2370.CrossRefGoogle ScholarPubMed
Dubois-Dalcq, M., Behar, T., Hudson, L., and Lazzarini, R. A. (1986). Emergence of three myelin proteins in oligodendrocytes cultured without neurons. J Cell Biol 102, 384–392.CrossRefGoogle ScholarPubMed
Duce, J. A., Hollander, W., Jaffe, R., and Abraham, C. R. (2006). Activation of early components of complement targets myelin and oligodendrocytes in the aged rhesus monkey brain. Neurobiol Aging 27, 633–644.CrossRefGoogle ScholarPubMed
Ducker, T. B., Salcman, M., Perot, P. L., and Ballantine, D. (1978). Experimental spinal cord trauma. I. Correlation of blood flow, tissue oxygen and neurologic status in the dog. Surg Neurol 10, 60–63.Google ScholarPubMed
Dugas, J. C., Tai, Y. C., Speed, T. P., Ngai, J., and Barres, B. A. (2006). Functional genomic analysis of oligodendrocyte differentiation. J Neurosci 26, 10967–10983.CrossRefGoogle ScholarPubMed
Duncan, I. D., Griffiths, I. R., and Munz, M. (1983). “Shaking pups”: a disorder of central myelination in the spaniel dog. III. Quantitative aspects of glia and myelin in the spinal cord and optic nerve. Neuropathol Appl Neurobiol 9, 355–368.CrossRefGoogle ScholarPubMed
Duncan, I. D., Hammang, J. P., and Trapp, B. D. (1988). Abnormal compact myelin in the myelin-deficient rat: absence of proteolipid protein correlates with a defect in the intraperiod line. Proc Natl Acad Sci USA 84, 6287–6291.CrossRefGoogle Scholar
Duncan, J. R., Cock, M. L., Scheerlinck, J. P., et al. (2002). White matter injury after repeated endotoxin exposure in the preterm ovine fetus. Pediatr Res 52, 941–949.CrossRefGoogle ScholarPubMed
Dupree, J. L., Coetzee, T., Blight, A., Suzuki, K., and Popko, B. (1998). Myelin galactolipids are essential for proper node of Ranvier formation in the CNS. J Neurosci 18, 1642–1649.CrossRefGoogle ScholarPubMed
Dupree, J. L., Girault, J. A., and Popko, B. (1999). Axo-glial interactions regulate the localization of axonal paranodal proteins. J Cell Biol 147, 1145–1152.CrossRefGoogle ScholarPubMed
Dyer, C. A. and Benjamins, J. A. (1989). Organization of oligodendroglial membrane sheets. I. Association of myelin basic protein and 2′,3′-cyclic nucleotide 3′-phosphohydrolase with cytoskeleton. J Neurosci Res 24, 201–211.CrossRefGoogle Scholar
Dzhashiashvili, Y., Zhang, Y., Galinska, J., Lam, I., Grumet, M., and Salzer, J. L. (2007). Nodes of Ranvier and axon initial segments are ankyrin G-dependent domains that assemble by distinct mechanisms. J Cell Biol 177, 857–870.CrossRefGoogle ScholarPubMed
Dziembowska, M., Tham, T. N., Lau, P., Vitry, S., Lazarini, F., and Dubois-Dalcq, M. (2005). A role for CXCR4 signaling in survival and migration of neural and oligodendrocyte precursors. Glia 50, 258–269.CrossRefGoogle ScholarPubMed
Eash, S., Tavares, R., Stopa, E. G., Robbins, S. H., Brossay, L., and Atwood, W. J. (2004). Differential distribution of the JC virus receptor-type sialic acid in normal human tissues. Am J Pathol 164, 419–428.CrossRefGoogle ScholarPubMed
Eberhardt, K. A., Irintchev, A., Al-Majed, A. A., et al. (2006). BDNF/TrkB signaling regulates HNK-1 carbohydrate expression in regenerating motor nerves and promotes functional recovery after peripheral nerve repair. Exp Neurol 198, 500–510.CrossRefGoogle ScholarPubMed
Edgar, J. M., McLaughlin, M., Yool, D., et al. (2004). Oligodendroglial modulation of fast axonal transport in a mouse model of hereditary spastic paraplegia. J Cell Biol 166, 121–131.CrossRefGoogle Scholar
Einheber, S., Zanazzi, G., Ching, W., et al. (1997). The axonal membrane protein Caspr, a homologue of neurexin IV, is a component of the septate-like paranodal junctions that assemble during myelination. J Cell Biol 139, 1495–1506.CrossRefGoogle ScholarPubMed
Einstein, O., Fainstein, N., Vaknin, I., et al. (2007). Neural precursors attenuate autoimmune encephalomyelitis by peripheral immunosuppression. Ann Neurol 61, 209–218.CrossRefGoogle ScholarPubMed
Einstein, O., Grigoriadis, N., Mizrachi-Kol, R., et al. (2006). Transplanted neural precursor cells reduce brain inflammation to attenuate chronic experimental autoimmune encephalomyelitis. Exp Neurol 198, 275–284.CrossRefGoogle ScholarPubMed
Eltayeb, S., Berg, A. L., Lassmann, H., et al. (2007). Temporal expression and cellular origin of CC chemokine receptors CCR1, CCR2 and CCR5 in the central nervous system: insight into mechanisms of MOG-induced EAE. J Neuroinflammation 4, 14.CrossRefGoogle ScholarPubMed
Engelhardt, B. (2008). The blood–central nervous system barriers actively control immune cell entry into the central nervous system. Curr Pharm Des 14, 1555–1565.CrossRefGoogle ScholarPubMed
Eshed, Y., Feinberg, K., Carey, D. J., and Peles, E. (2007). Secreted gliomedin is a perinodal matrix component of peripheral nerves. J Cell Biol 177, 551–562.CrossRefGoogle ScholarPubMed
Eshed, Y., Feinberg, K., Poliak, S., et al. (2005). Gliomedin mediates Schwann cell–axon interaction and the molecular assembly of the nodes of Ranvier. Neuron 47, 215–229.CrossRefGoogle ScholarPubMed
Esper, R. M., Pankonin, M. S., and Loeb, J. A. (2006). Neuregulins: versatile growth and differentiation factors in nervous system development and human disease. Brain Res Rev 51, 161–175.CrossRefGoogle ScholarPubMed
Fairman, W. A., Vandenberg, R. J., Arriza, J. L., Kavanaugh, M. P., and Amara, S. G. (1995). An excitatory amino-acid transporter with properties of a ligand-gated chloride channel. Nature 375, 599–603.CrossRefGoogle ScholarPubMed
Falls, D. L. (2003). Neuregulins: functions, forms, and signaling strategies. Exp Cell Res 284, 14–30.CrossRefGoogle ScholarPubMed
Fancy, S. P., Zhao, C., and Franklin, R. J. (2004). Increased expression of Nkx2.2 and Olig2 identifies reactive oligodendrocyte progenitor cells responding to demyelination in the adult CNS. Mol Cell Neurosci 27, 247–254.CrossRefGoogle ScholarPubMed
Fannon, A. M., Sherman, D. L., Ilyina-Gragerova, G., Brophy, P. J., Friedrich, V. L., and Colman, D. R. (1995). Novel E-cadherin-mediated adhesion in peripheral nerve: Schwann cell architecture is stabilized by autotypic adherens junctions. J Cell Biol 129, 189–202.CrossRefGoogle ScholarPubMed
Farh, K. K., Grimson, A., Jan, C., et al. (2005). The widespread impact of mammalian MicroRNAs on mRNA repression and evolution. Science 310, 1817–1821.CrossRefGoogle ScholarPubMed
Fehlings, M. G. and Nashmi, R. (1995). Assessment of axonal dysfunction in an in vitro model of acute compressive injury to adult rat spinal cord axons. Brain Res 677, 291–299.CrossRefGoogle Scholar
Fehlings, M. G. and Tator, C. H. (1995). The relationships among the severity of spinal cord injury, residual neurological function, axon counts, and counts of retrogradely labeled neurons after experimental spinal cord injury. Exp Neurol 132, 220–228.CrossRefGoogle ScholarPubMed
Feldman, M. L., and Peters, A. (1998). Ballooning of myelin sheaths in normally aged macaques. J Neurocytol 27, 605–614.CrossRefGoogle ScholarPubMed
Feltri, M. L., Graus Porta, D., Previtali, S. C., et al. (2002). Conditional disruption of beta 1 integrin in Schwann cells impedes interactions with axons. J Cell Biol 156, 199–209.CrossRefGoogle Scholar
Ferguson, B., Matyszak, M. K., Esiri, M. M., and Perry, V. H. (1997). Axonal damage in acute multiple sclerosis lesions. Brain 120, 393–399.CrossRefGoogle ScholarPubMed
Fern, R. and Möller, T. (2000). Rapid ischemic cell death in immature oligodendrocytes: a fatal glutamate release feedback loop. J Neurosci 20, 34–42.CrossRefGoogle ScholarPubMed
Fernandez-Valle, C., Bunge, R. P., and Bunge, M. B. (1995). Schwann cells degrade myelin and proliferate in the absence of macrophages: evidence from in vitro studies of Wallerian degeneration. J Neurocytol 24, 667–679.CrossRefGoogle ScholarPubMed
Ferriero, D. M. (2006). Can we define the pathogenesis of human periventricular white-matter injury using animal models? J Child Neurol 21, 580–581.CrossRefGoogle ScholarPubMed
Fewou, S. N., Ramakrishnan, H., Bussow, H., Gieselmann, V., and Eckhardt, M. (2007). Down-regulation of polysialic acid is required for efficient myelin formation. J Biol Chem 282, 16700–16711.CrossRefGoogle ScholarPubMed
Ffrench-Constant, C., and Raff, M. C. (1986). Proliferating bipotential glial progenitor cells in adult rat optic nerve. Nature 319, 499–502.CrossRefGoogle ScholarPubMed
Fields, R. D. (2005). Myelination: an overlooked mechanism of synaptic plasticity? Neuroscientist 11, 528–531.CrossRefGoogle ScholarPubMed
Fields, R. D. (2008a). Oligodendrocytes changing the rules: action potentials in glia and oligodendrocytes controlling action potentials. Neuroscientist 14, 540–543.CrossRefGoogle ScholarPubMed
Fields, R. D. (2008b). White matter in learning, cognition and psychiatric disorders. Trends Neurosci 31, 361–370.CrossRefGoogle ScholarPubMed
Fields, R. D. (2008c). White matter matters. Sci Am 298, 42–49.Google ScholarPubMed
Fields, R. D. and Burnstock, G. (2006). Purinergic signalling in neuron-glia interactions. Nat Rev Neurosci 7, 423–436.CrossRefGoogle ScholarPubMed
Filbin, M. T. (2003). Myelin-associated inhibitors of axonal regeneration in the adult mammalian CNS. Nat Rev Neurosci 4, 703–713.CrossRefGoogle ScholarPubMed
Filippi, M. and Rocca, M. A. (2005). MRI evidence for multiple sclerosis as a diffuse disease of the central nervous system. J Neurol 252 Suppl 5, v16–v24.CrossRefGoogle ScholarPubMed
Fischer, D., He, Z., and Benowitz, L. I. (2004). Counteracting the Nogo receptor enhances optic nerve regeneration if retinal ganglion cells are in an active growth state. J Neurosci 24, 1646–1651.CrossRefGoogle Scholar
Fischer, I., Konola, J., and Cochary, E. (1990). Microtubule associated protein (MAP1B) is present in cultured oligodendrocytes and co-localizes with tubulin. J Neurosci Res 27, 112–124.CrossRefGoogle ScholarPubMed
Fix, A. S., Horn, J. W., Wightman, K. A., et al. (1993). Neuronal vacuolization and necrosis induced by the noncompetitive N-methyl-d-aspartate (NMDA) antagonist MK(+)801 (dizocilpine maleate): a light and electron microscopic evaluation of the rat retrosplenial cortex. Exp Neurol 123, 204–215.CrossRefGoogle ScholarPubMed
Flores, A. I., Narayanan, S. P., Morse, E. N., et al. (2008). Constitutively active Akt induces enhanced myelination in the CNS. J Neurosci 28, 7174–7183.CrossRefGoogle ScholarPubMed
Fogarty, M., Richardson, W. D., and Kessaris, N. (2005). A subset of oligodendrocytes generated from radial glia in the dorsal spinal cord. Development 132, 1951–1959.CrossRefGoogle ScholarPubMed
Folkerth, R. D. and Kinney, H. C. (2008). Disorders of the perinatal period. In Greenfield's Neuropathology (London: Hodder Arnold), pp. 241–315.Google Scholar
Follett, P. L., Rosenberg, P. A., Volpe, J. J., and Jensen, F. E. (2000). NBQX attenuates excitotoxic injury in developing white matter. J Neurosci 20, 9235–9241.CrossRefGoogle ScholarPubMed
Follett, P. L., Talos, D. M., Volpe, J. J., and Jensen, F. E. (2003). AMPA receptor subunit expression in human white matter during window of susceptibility to periventricular leukomalacia. Ann Neurol 54(S3), S103.Google Scholar
Foote, A. K. and Blakemore, W. F. (2005). Inflammation stimulates remyelination in areas of chronic demyelination. Brain 128, 528–539.CrossRefGoogle ScholarPubMed
Fortin, D., Rom, E., Sun, H., Yayon, A., and Bansal, R. (2005). Distinct fibroblast growth factor (FGF)/FGF receptor signaling pairs initiate diverse cellular responses in the oligodendrocyte lineage. J Neurosci 25, 7470–7479.CrossRefGoogle ScholarPubMed
Fournier, A. E., GrandPre, T., and Strittmatter, S. M. (2001). Identification of a receptor mediating Nogo-66 inhibition of axonal regeneration. Nature 409, 341–346.CrossRefGoogle ScholarPubMed
Fournier, A. E., Takizawa, B. T., and Strittmatter, S. M. (2003). Rho kinase inhibition enhances axonal regeneration in the injured CNS. J Neurosci 23, 1416–1423.CrossRefGoogle ScholarPubMed
Fox, E. J. (2004). Mechanism of action of mitoxantrone. Neurology 63, S15–S18.CrossRefGoogle ScholarPubMed
Franco, P. G., Silvestroff, L., Soto, E. F., and Pasquini, J. M. (2008). Thyroid hormones promote differentiation of oligodendrocyte progenitor cells and improve remyelination after cuprizone-induced demyelination. Exp Neurol 212, 458–467.CrossRefGoogle ScholarPubMed
Frank, M. (2000). MAL, a proteolipid in glycosphingolipid enriched domains: functional implications in myelin and beyond. Prog Neurobiol 60, 531–544.CrossRefGoogle ScholarPubMed
Franklin, R. J. (2002). Why does remyelination fail in multiple sclerosis? Nat Rev Neurosci 3, 705–714.CrossRefGoogle ScholarPubMed
Franklin, R. J. and Barnett, S. C. (2004). Olfactory ensheathing cells. In Myelin Biology and Disorders, Lazzarini, R. A., ed. (New York: Elsevier), pp. 371–384.Google Scholar
Franklin, R. J. and Ffrench-Constant, C. (2008). Remyelination in the CNS: from biology to therapy. Nat Rev Neurosci 9, 839–855.CrossRefGoogle Scholar
Freedman, M. S., Buu, N. N., Ruijs, T. C., Williams, K., and Antel, J. P. (1992). Differential expression of heat shock proteins by human glial cells. J Neuroimmunol 41, 231–238.CrossRefGoogle ScholarPubMed
Freedman, M. S., Ruijs, T. C., Selin, L. K., and Antel, J. P. (1991). Peripheral blood gamma-delta T cells lyse fresh human brain-derived oligodendrocytes. Ann Neurol 30, 794–800.CrossRefGoogle ScholarPubMed
Friede, R. L. (1972). Control of myelin formation by axon caliber (with a model of the control mechanism). J Comp Neurol 144, 233–252.CrossRefGoogle Scholar
Frost, E. E., Zhou, Z., Krasnesky, K., and Armstrong, R. C. (2009). Initiation of oligodendrocyte progenitor cell migration by a PDGF-A activated extracellular regulated kinase (ERK) signaling pathway. Neurochem Res 34, 169–181.CrossRefGoogle ScholarPubMed
Fruttiger, M., Montag, D., Schachner, M., and Martini, R. (1995). Crucial role for the myelin-associated glycoprotein in the maintenance of axon-myelin integrity. Eur J Neurosci 7, 511–515.CrossRefGoogle ScholarPubMed
Fry, E. J., Ho, C., and David, S. (2007). A role for Nogo receptor in macrophage clearance from injured peripheral nerve. Neuron 53, 649–662.CrossRefGoogle ScholarPubMed
Fukushima, N., Furuta, D., Hidaka, Y., Moriyama, R., and Tsujiuchi, T. (2009). Post-translational modifications of tubulin in the nervous system. J Neurochem 109, 683–693.CrossRefGoogle Scholar
Furley, A. J., Morton, S. B., Manalo, D., Karagogeos, D., Dodd, J., and Jessell, T. M. (1990). The axonal glycoprotein TAG-1 is an immunoglobulin superfamily member with neurite outgrowth-promoting activity. Cell 61, 157–170.CrossRefGoogle ScholarPubMed
Furness, D. N., Dehnes, Y., Akhtar, A. Q., et al. (2008). A quantitative assessment of glutamate uptake into hippocampal synaptic terminals and astrocytes: new insights into a neuronal role for excitatory amino acid transporter 2 (EAAT2). Neuroscience 157, 80–94.CrossRefGoogle Scholar
Furuta, A., Rothstein, J. D., and Martin, L. J. (1997). Glutamate transporter protein subtypes are expressed differentially during rat CNS development. J Neurosci 17, 8363–8375.CrossRefGoogle ScholarPubMed
Furuta, A., Takashima, S., Yokoo, H., Rothstein, J. D., Wada, K., and Iwaki, T. (2005). Expression of glutamate transporter subtypes during normal human corticogenesis and type II lissencephaly. Brain Res Dev Brain Res 155, 155–164.CrossRefGoogle ScholarPubMed
Gage, F. H. (2000). Mammalian neural stem cells. Science 287, 1433–1438.CrossRefGoogle ScholarPubMed
Galiano, M. R., Bosc, C., Schweitzer, A., Andrieux, A., Job, D., and Hallak, M. E. (2004). Astrocytes and oligodendrocytes express different STOP protein isoforms. J Neurosci Res 78, 329–337.CrossRefGoogle ScholarPubMed
Gallo, V. and Russell, J. T. (1995). Excitatory amino acid receptors in glia: different subtypes for distinct functions. J. Neurosci. Res. 42, 1–8.CrossRefGoogle ScholarPubMed
Gallo, V., Mangin, J. M., Kukley, M., and Dietrich, D. (2008). Synapses on NG2-expressing progenitors in the brain: multiple functions? J Physiol (Lond) 586, 3767–3781.CrossRefGoogle ScholarPubMed
Gallo, V., Zhou, J. M., McBain, C. J., Wright, P., Knutson, P. L., and Armstrong, R. C. (1996). Oligodendrocyte progenitor cell proliferation and lineage progression are regulated by glutamate receptor-mediated K+ channel block. J Neurosci 16, 2659–2670.CrossRefGoogle ScholarPubMed
Gao, L. and Miller, R. H. (2006). Specification of optic nerve oligodendrocyte precursors by retinal ganglion cell axons. J Neurosci 29, 7619–7628.CrossRefGoogle Scholar
Gao, X., Gillig, T. A., Ye, P., D'Ercole, A. J., Matsushima, G. K., and Popko, B. (2000). Interferon-gamma protects against cuprizone-induced demyelination. Mol Cell Neurosci 16, 338–349.CrossRefGoogle ScholarPubMed
Garbern, J. Y. (2007). Pelizaeus–Merzbacher disease: genetic and cellular pathogenesis. Cell Mol Life Sci 64, 50–65.CrossRefGoogle ScholarPubMed
Garbern, J. Y., Cambi, F., Tang, X. M., et al. (1997). Proteolipid protein is necessary in peripheral as well as central myelin. Neuron 19, 205–218.CrossRefGoogle ScholarPubMed
Garbern, J. Y., Yool, D. A., Moore, G. J., et al. (2002). Patients lacking the major CNS myelin protein, proteolipid protein 1, develop length-dependent axonal degeneration in the absence of demyelination and inflammation. Brain 125, 551–561.CrossRefGoogle ScholarPubMed
Garcia, R., Aguiar, J., Alberti, E., Cuetara, K., and Pavon, N. (2004). Bone marrow stromal cells produce nerve growth factor and glial cell line-derived neurotrophic factors. Biochem Biophys Res Commun 316, 753–754.CrossRefGoogle ScholarPubMed
Garcia-Barcina, J. M. and Matute, C. (1996). Expression of kainate-selective glutamate receptor subunits in glial cells of the adult bovine white matter. Eur J Neurosci 8, 2379–2387.CrossRefGoogle ScholarPubMed
Gard, A. L., Williams, W. C., and Burrell, M. R. (1995). Oligodendroblasts distinguished from O-2A glial progenitors by surface phenotype (O4+GalC–) and response to cytokines using signal transducer LIFR beta. Dev Biol 167, 596–608.CrossRefGoogle ScholarPubMed
Garlin, A. B., Sinor, A. D., Sinor, J. D., Jee, S. H., Grinspan, J. B., and Robinson, M. B. (1995). Pharmacology of sodium-dependent high-affinity l-[3H]glutamate transport in glial cultures. J Neurochem 64, 2572–2580.CrossRefGoogle ScholarPubMed
Garratt, A. N., Britsch, S., and Birchmeier, C. (2000). Neuregulin, a factor with many functions in the life of a Schwann cell. Bioessays 22, 987–996.3.0.CO;2-5>CrossRefGoogle ScholarPubMed
Gasque, P. and Morgan, B. P. (1996). Complement regulatory protein expression by a human oligodendrocyte cell line: cytokine regulation and comparison with astrocytes. Immunology 89, 338–347.CrossRefGoogle ScholarPubMed
Gasque, P., Singhrao, S. K., Neal, J. W., Gotze, O., and Morgan, B. P. (1997). Expression of the receptor for complement C5a (CD88) is up-regulated on reactive astrocytes, microglia, and endothelial cells in the inflamed human central nervous system. Am J Pathol 150, 31–41.Google ScholarPubMed
Gatzinsky, K. P., Persson, G. H., and Berthold, C. H. (1997). Removal of retrogradely transported material from rat lumbosacral alpha-motor axons by paranodal axon–Schwann cell networks. Glia 20, 115–126.3.0.CO;2-8>CrossRefGoogle ScholarPubMed
Gay, C. T., Hardies, L. J., Rauch, R. A., et al. (1997). Magnetic resonance imaging demonstrates incomplete myelination in 18q- syndrome: evidence for myelin basic protein haploinsufficiency. Am J Med Genet 74, 422–431.3.0.CO;2-K>CrossRefGoogle ScholarPubMed
Ge, W. P., Zhou, W., Luo, Q., Jan, L. Y., and Jan, Y. N. (2009). Dividing glial cells maintain differentiated properties including complex morphology and functional synapses. Proc Natl Acad Sci USA 106, 328–333.CrossRefGoogle ScholarPubMed
Geiger, J. R., Melcher, T., Koh, D. S., et al. (1995). Relative abundance of subunit mRNAs determines gating and Ca2+ permeability of AMPA receptors in principal neurons and interneurons in rat CNS. Neuron 15, 193–204.CrossRefGoogle ScholarPubMed
Geiger, K., Howes, E., Gallina, M., Huang, X. J., Travis, G. H., and Sarvetnick, N. (1994). Transgenic mice expressing IFN-gamma in the retina develop inflammation of the eye and photoreceptor loss. Invest Ophthalmol Vis Sci 35, 2667–2681.Google ScholarPubMed
Genain, C. P., Cannella, B., Hauser, S. L., and Raine, C. S. (1999). Identification of autoantibodies associated with myelin damage in multiple sclerosis. Nature Med 5, 170–175.CrossRefGoogle ScholarPubMed
Genain, C. P., Nguyen, M. H., Letvin, N. L., et al. (1995). Antibody facilitation of multiple sclerosis-like lesions in a nonhuman primate. J Clin Invest 96, 2966–2974.CrossRefGoogle Scholar
Gencic, S. and Hudson, L. D. (1990). Conservative amino acid substitution in the myelin proteolipid protein of jimpymsd mice. J Neurosci 10, 117–124.CrossRefGoogle ScholarPubMed
Genoud, S., Lappe-Siefke, C., Goebbels, S., et al. (2002). Notch1 control of oligodendrocyte differentiation in the spinal cord. J Cell Biol 158, 709–718.CrossRefGoogle ScholarPubMed
Gensert, J. M. and Goldman, J. E. (1997). Endogenous progenitors remyelinate demyelinated axons in the adult CNS. Neuron 19, 197–203.CrossRefGoogle ScholarPubMed
George, E. B., Glass, J. D., and Griffin, J. W. (1995). Axotomy-induced axonal degeneration is mediated by calcium influx through ion-specific channels. J Neurosci 15, 6445–6452.CrossRefGoogle ScholarPubMed
Gerdoni, E., Gallo, B., Casazza, S., et al. (2007). Mesenchymal stem cells effectively modulate pathogenic immune response in experimental autoimmune encephalomyelitis. Ann Neurol 61, 219–227.CrossRefGoogle ScholarPubMed
Ghabriel, M. N. and Allt, G. (1981). Incisures of Schmidt–Lanterman. Prog Neurobiol 17, 25–58.CrossRefGoogle ScholarPubMed
Ghandour, M. S. and Skoff, R. P. (1991). Double-labeling in situ hybridization analysis of mRNAs for carbonic anhydrase II and myelin basic protein: expression in developing cultured glial cells. Glia 4, 1–10.CrossRefGoogle ScholarPubMed
Ghatak, N. R., Hirano, A., Doron, Y., and Zimmerman, H. M. (1973). Remyelination in multiple sclerosis with peripheral type myelin. Arch Neurol 29, 262–267.CrossRefGoogle ScholarPubMed
Giese, K. P., Martini, R., Lemke, G., Soriano, P., and Schachner, M. (1992). Mouse P0 gene disruption leads to hypomyelination, abnormal expression of recognition molecules, and degeneration of myelin and axons. Cell 71, 565–576.CrossRefGoogle ScholarPubMed
Gill, A. S. and Binder, D. K. (2007). Wilder Penfield, Pio del Hortega, and the discovery of oligodendroglia. Neurosurgey 60, 940–948.CrossRefGoogle ScholarPubMed
Gilles, F. H., Leviton, A., and Kerr, C. S. (1976). Endotoxin leucoencephalopathy in the telencephalon of the newborn kitten. J Neurol Sci 27, 183–191.CrossRefGoogle ScholarPubMed
Gillespie, C. S., Trapp, B. D., Colman, D. R., and Brophy, P. J. (1990). Distribution of myelin basic protein and P2 mRNA's in rabbit spinal cord oligodendrocytes. J Neurochem 54, 1556–1561.CrossRefGoogle Scholar
Gilmour, D. T., Maischein, H. M., and Nusslein-Volhard, C. (2002). Migration and function of a glial subtype in the vertebrate peripheral nervous system. Neuron 34, 577–588.CrossRefGoogle ScholarPubMed
Giovannoni, G. and Hartung, H. P. (1996). The immunopathogenesis of multiple sclerosis and Guillain–Barre syndrome. Curr Opin Neurol 9, 165–177.CrossRefGoogle ScholarPubMed
Giraldez, A. J., Mishima, Y., Rihel, J., et al. (2006). Zebrafish MiR-430 promotes deadenylation and clearance of maternal mRNAs. Science 312, 75–79.CrossRefGoogle ScholarPubMed
Giulian, D., Young, D. G., Woodward, J., Brown, D. C., and Lachman, L. B. (1988). Interleukin-1 is an astroglial growth factor in the developing brain. J Neurosci 8, 709–714.CrossRefGoogle ScholarPubMed
Giuliani, F., Goodyer, C. G., Antel, J. P., and Yong, V. W. (2003). Vulnerability of human neurons to T cell-mediated cytotoxicity. J Immunol 171, 368–379.CrossRefGoogle ScholarPubMed
Gledhill, R. F. and McDonald, W. I. (1977). Morphological characteristics of central demyelination and remyelination: a single-fiber study. Ann Neurol 1, 552–560.CrossRefGoogle ScholarPubMed
Gledhill, R. F.Harrison, B. M., and McDonald, W. I. (1973a). Demyelination and remyelination after acute spinal cord compression. Exp Neurol 38, 472–487.CrossRefGoogle ScholarPubMed
Gledhill, R. F., Harrison, B. M., and McDonald, W. I. (1973b). Pattern of remyelination in the CNS. Nature 244, 443–444.CrossRefGoogle ScholarPubMed
Gocht, A. (1992). The subcellular localization of the carbohydrate epitope 3-fucosyl-N-acetyllactosamine is different in normal and reactive astrocytes. Acta Anat (Basel) 145, 434–441.CrossRefGoogle ScholarPubMed
Goddard, D. R., Berry, M., and Butt, A. M. (1999). In vivo actions of fibroblast growth factor-2 and insulin-like growth factor-I on oligodendrocyte development and myelination in the central nervous system. J Neurosci Res 57, 74–85.3.0.CO;2-O>CrossRefGoogle ScholarPubMed
Goldberg, J. L., Vargas, M. E., Wang, J. T., et al. (2004). An oligodendrocyte lineage-specific semaphorin, Sema5A, inhibits axon growth by retinal ganglion cells. J Neurosci 24, 4989–4999.CrossRefGoogle ScholarPubMed
Goldman, J. E., Geier, S. S., and Hirano, M. (1986). Differentiation of astrocytes and oligodendrocytes from germinal matrix cells in primary culture. J Neurosci 6, 52–60.CrossRefGoogle ScholarPubMed
Goldstein, L. S. and Yang, Z. (2000). Microtubule-based transport systems in neurons: the roles of kinesins and dyneins. Annu Rev Neurosci 23, 39–71.CrossRefGoogle ScholarPubMed
Gollan, L., Sabanay, H., Poliak, S., Berglund, E. O., Ranscht, B., and Peles, E. (2002). Retention of a cell adhesion complex at the paranodal junction requires the cytoplasmic region of Caspr. J Cell Biol 157, 1247–1256.CrossRefGoogle ScholarPubMed
Gomes, W. A., Mehler, M. F., and Kessler, J. A. (2003). Transgenic overexpression of BMP4 increases astroglial and decreases oligodendroglial lineage commitment. Dev Biol 255, 164–177.CrossRefGoogle ScholarPubMed
Gorath, M., Stahnke, T., Mronga, T., Goldbaum, O., and Richter-Landsberg, C. (2001). Developmental changes of tau protein and mRNA in cultured rat brain oligodendrocytes. Glia 36, 89–101.CrossRefGoogle ScholarPubMed
Gordon, D., Kidd, G. J., and Smith, R. (2008a). Antisense suppression of tau in cultured rat oligodendrocytes inhibits process formation. J Neurosci Res 86, 2591–2601.CrossRefGoogle ScholarPubMed
Gordon, D., Pavlovska, G., Glover, C. P., Uney, J. B., Wraith, D., and Scolding, N. J. (2008b). Human mesenchymal stem cells abrogate experimental allergic encephalomyelitis after intraperitoneal injection, and with sparse CNS infiltration. Neurosci Lett 448, 71–73.CrossRefGoogle ScholarPubMed
Gorman, M. P., Golomb, M. R., Walsh, L. E., et al. (2007). Steroid-responsive neurologic relapses in a child with a proteolipid protein-1 mutation. Neurology 68, 1305–1307.CrossRefGoogle Scholar
Gould, R. M., Byrd, A. L., and Barbarese, E. (1995). The number of Schmidt–Lanterman incisures is more than doubled in shiverer PNS myelin sheaths. J Neurocytol 24, 85–98.CrossRefGoogle ScholarPubMed
Gould, R. M., Freund, C. M., Palmer, F., and Feinstein, D. L. (2000). Messenger RNAs located in myelin sheath assembly sites. J Neurochem 75, 1834–1844.CrossRefGoogle ScholarPubMed
Gow, A. and Lazzarini, R. A. (1996). A cellular mechanism governing the severity of Pelizaeus–-Merzbacher disease. Nat Genet 13, 422–428.CrossRefGoogle ScholarPubMed
Gow, A., Friedrich, V. L., and Lazzarini, R. A. (1994). Many naturally occurring mutations of myelin proteolipid protein impair its intracellular transport. J Neursci Res 37, 574–583.CrossRefGoogle ScholarPubMed
Gow, A., Southwood, C. M., and Lazzarini, R. A. (1998). Disrupted proteolipid protein trafficking results in oligodendrocyte apoptosis in an animal model of Pelizaeus–Merzbacher disease. J Cell Biol 140, 925–934.CrossRefGoogle Scholar
Gow, A., Southwood, C. M., Li, J. S., et al. (1999). CNS myelin and Sertoli cell tight junction strands are absent in Osp/claudin-11 null mice. Cell 99, 649–659.CrossRefGoogle ScholarPubMed
GrandPre, T., Li, S., and Strittmatter, S. M. (2002). Nogo-66 receptor antagonist peptide promotes axonal regeneration. Nature 417, 547–551.CrossRefGoogle ScholarPubMed
GrandPre, T., Nakamura, F., Vartanian, T., and Strittmatter, S. M. (2000). Identification of the Nogo inhibitor of axon regeneration as a Reticulon protein. Nature 403, 439–444.CrossRefGoogle ScholarPubMed
Graumann, U., Reynolds, R., Steck, A. J., and Schaeren-Wiemers, N. (2003). Molecular changes in normal appearing white matter in multiple sclerosis are characteristic of neuroprotective mechanisms against hypoxic insult. Brain Pathol 13, 554–573.CrossRefGoogle ScholarPubMed
Gregori, N., Proschel, C., Noble, M., and Mayer-Proschel, M. (2002). The tripotential glial-restricted precursor (GRP) cell and glial development in the spinal cord: generation of bipotential oligodendrocyte-type-2 astrocyte progenitor cells and dorsal-ventral differences in GRP cell function. J Neurosci 22, 248–256.CrossRefGoogle ScholarPubMed
Grenier, Y., Ruijs, T. C., Robitaille, Y., Olivier, A., and Antel, J. P. (1989). Immunohistochemical studies of adult human glial cells. J Neuroimmunol 21, 103–115.CrossRefGoogle ScholarPubMed
Gressens, P., Marret, S., and Evrard, P. (1996). Developmental spectrum of the excitotoxic cascade induced by ibotenate: a model of hypoxic insults in fetuses and neonates. Neuropathol Appl Neurobiol 22, 498–502.CrossRefGoogle ScholarPubMed
Griffiths, I. and McCulloch, M. C. (1983). Nerve fibres in spinal cord impact injuries. Part 1. Changes in the myelin sheath during the initial 5 weeks. J Neurol Sci 58, 335–349.CrossRefGoogle ScholarPubMed
Griffiths, I., Klugmann, M., Anderson, T., Thomson, C., Vouyiouklis, D., and Nave, K. A. (1998a). Current concepts of PLP and its role in the nervous system. Microsc Res Tech 41, 344–358.3.0.CO;2-Q>CrossRefGoogle Scholar
Griffiths, I., Klugmann, M., Anderson, T., et al. (1998b). Axonal swellings and degeneration in mice lacking the major proteolipid of myelin. Science 280, 1610–1613.CrossRefGoogle ScholarPubMed
Grinspan, J. B., Stern, J. L., Franceschini, B., and Pleasure, D. (1993). Trophic effects of basic fibroblast growth factor (bFGF) on differentiated oligodendroglia: a mechanism for regeneration of the oligodendroglial lineage. J Neurosci Res 36, 672–680.CrossRefGoogle ScholarPubMed
Griot-Wenk, M., Griot, C., Pfister, H., and Vandevelde, M. (1991). Antibody-dependent cellular cytotoxicity in antimyelin antibody-induced oligodendrocyte damage in vitro. J Neuroimmunol 33, 145–155.CrossRefGoogle ScholarPubMed
Groom, A. J., Smith, T., and Turski, L. (2003). Multiple sclerosis and glutamate. Ann N Y Acad Sci 993, 229–275.CrossRefGoogle ScholarPubMed
Grossman, S. D., Rosenberg, L. J., and Wrathall, J. R. (2001). Temporal-spatial pattern of acute neuronal and glial loss after spinal cord contusion. Exp Neurol 168, 273–282.CrossRefGoogle ScholarPubMed
Grove, M., Komiyama, N. H., Nave, K. A., Grant, S. G., Sherman, D. L., and Brophy, P. J. (2007). FAK is required for axonal sorting by Schwann cells. J Cell Biol 176, 277–282.CrossRefGoogle ScholarPubMed
Groves, A. K., Barnett, S. C., Franklin, R. J. M., et al. (1993). Repair of demyelinated lesions by transplantation of purified O-2A progenitor cells. Nature 362, 453–455.CrossRefGoogle ScholarPubMed
Grumet, M., Mauro, V., Burgoon, M. P., Edelman, G. M., and Cunningham, B. A. (1991). Structure of a new nervous system glycoprotein, Nr-CAM, and its relationship to subgroups of neural cell adhesion molecules. J Cell Biol 113(6), 1399–1412.CrossRefGoogle ScholarPubMed
Gudz, T. I., Komuro, H., and Macklin, W. B. (2006). Glutamate stimulates oligodendrocyte progenitor migration mediated via an alphav integrin/myelin proteolipid protein complex. J Neurosci 26, 2458–2466.CrossRefGoogle ScholarPubMed
Guertin, A. D., Zhang, D. P., Mak, K. S., Alberta, J. A., and Kim, H. A. (2005). Microanatomy of axon/glial signaling during Wallerian degeneration. J Neurosci 25, 3478–3487.CrossRefGoogle ScholarPubMed
Guidotti, G. (1972). Membrane proteins. Annu Rev Biochem 41, 731–752.CrossRefGoogle ScholarPubMed
Hafler, D. A., Compston, A., Sawcer, S., et al. (2007). Risk alleles for multiple sclerosis identified by a genomewide study. N Engl J Med 357, 851–862.Google ScholarPubMed
Hagberg, H. (1992). Hypoxic-ischemic damage in the neonatal brain: excitatory amino acids. Dev Pharmacol Ther 18, 139–144.Google ScholarPubMed
Haines, J. D., Fragoso, G., Hossain, S., Mushynski, W. E., and Almazan, G. (2008). p38 mitogen-activated protein kinase regulates myelination. J Mol Neurosci 35, 23–33.CrossRefGoogle ScholarPubMed
Hakak, Y., Walker, J. R., Li, C., et al. (2001). Genome-wide expression analysis reveals dysregulation of myelination-related genes in chronic schizophrenia. Proc Natl Acad Sci USA 98, 4746–4751.CrossRefGoogle ScholarPubMed
Halfpenny, C. A. and Scolding, N. J. (2003). Immune-modifying agents do not impair the survival, migration or proliferation of oligodendrocyte progenitors (CG-4) in vitro. J. Neuroimmunol. 139, 9–16.CrossRefGoogle ScholarPubMed
Halfpenny, C., Benn, T., and Scolding, N. J. (2002). Cell transplantation, myelin repair and multiple sclerosis. Lancet Neurol 1, 31–40.CrossRefGoogle ScholarPubMed
Hall, A. K. and Miller, R. (2004). Emerging roles for bone morphogenetic proteins in central nervous system glial biology. J Neurosci Res 76, 1–8.CrossRefGoogle ScholarPubMed
Hall, S. (2005). The response to injury in the peripheral nervous system. J Bone Joint Surg Br 87, 1309–1319.CrossRefGoogle ScholarPubMed
Hall, S. M. and Gregson, N. A. (1974). The effects of mitomycin C on remyelination in the peripheral nervous system. Nature 252, 303–305.CrossRefGoogle ScholarPubMed
Hamamoto, Y., Ogata, T., Morino, T., Hino, M., and Yamamoto, H. (2007). Real-time direct measurement of spinal cord blood flow at the site of compression: relationship between blood flow recovery and motor deficiency in spinal cord injury. Spine 32, 1955–1962.CrossRefGoogle ScholarPubMed
Hamanoue, M., Yoshioka, A., Ohashi, T., Eto, Y., and Takamatsu, K. (2004). NF-kappaB prevents TNF-alpha-induced apoptosis in an oligodendrocyte cell line. Neurochem Res 29, 1571–1576.CrossRefGoogle Scholar
Hamrick, S. E., Miller, S. P., Leonard, C., et al. (2004). Trends in severe brain injury and neurodevelopmental outcome in premature newborn infants: the role of cystic periventricular leukomalacia. J Pediatr 145, 593–599.CrossRefGoogle ScholarPubMed
Haney, C. A., Sahenk, Z., Li, C., Lemmon, V. P., Roder, J., and Trapp, B. D. (1999). Heterophilic binding of L1 on unmyelinated sensory axons mediates Schwann cell adhesion and is required for axonal survival. J Cell Biol 146, 1173–1184.CrossRefGoogle ScholarPubMed
Hansen, T., Olsen, L., Lindow, M., et al. (2007). Brain expressed microRNAs implicated in schizophrenia etiology. PLoS ONE 2, e873.CrossRefGoogle ScholarPubMed
Hansen-Pupp, I., Harling, S., Berg, A. C., Cilio, C., Hellstrom-Westas, L., and Ley, D. (2005). Circulating interferon-gamma and white matter brain damage in preterm infants. Pediatr Res 58, 946–952.CrossRefGoogle ScholarPubMed
Harada, A., Oguchi, K., Okabe, S., et al. (1994). Altered microtubule organization in small-calibre axons of mice lacking tau protein. Nature 369, 488–491.CrossRefGoogle ScholarPubMed
Harauz, G., Ishiyama, N., Hill, C. M., Bates, I. R., Libich, D. S., and Fares, C. (2004). Myelin basic protein-diverse conformational states of an intrinsically unstructured protein and its roles in myelin assembly and multiple sclerosis. Micron 35, 503–542.CrossRefGoogle ScholarPubMed
Harel, N. Y. and Strittmatter, S. M. (2006). Can regenerating axons recapitulate developmental guidance during recovery from spinal cord injury? Nat Rev Neurosci 7, 603–616.CrossRefGoogle ScholarPubMed
Harrison, B. M. and McDonald, W. I. (1977). Remyelination after transient experimental compression of the spinal cord. Ann Neurol 1, 542–551.CrossRefGoogle ScholarPubMed
Harrison, B. M., Gledhill, R. F., and McDonald, W. J. (1975). Remyelination after transient compression of the spinal cord. Proc Aust Assoc Neurol 12, 117–122.Google ScholarPubMed
Harsan, L. A., Steibel, J., Zaremba, A., et al. (2008). Recovery from chronic demyelination by thyroid hormone therapy: myelinogenesis induction and assessment by diffusion tensor magnetic resonance imaging. J Neurosci 28, 14189–14201.CrossRefGoogle ScholarPubMed
Hart, I. K., Richardson, W. D., Heldin, C. H., Westermark, B., and Raff, M. C. (1989). PDGF receptors on cells of the oligodendrocyte-type-2 astrocyte (O-2A) cell lineage. Development 105, 595–603.Google ScholarPubMed
Hartline, D. K. and Colman, D. R. (2007). Rapid conduction and the evolution of giant axons and myelinated fibers. Curr Biol 17, R29–R35.CrossRefGoogle ScholarPubMed
Hartung, H. P. (2009). New cases of progressive multifocal leukoencephalopathy after treatment with natalizumab. Lancet Neurol 8, 28–31.CrossRefGoogle ScholarPubMed
Hartung, H. P., Gonsette, R., Konig, N., et al. (2002). Mitoxantrone in progressive multiple sclerosis: a placebo-controlled, double-blind, randomised, multicentre trial. Lancet 360, 2018–2025.CrossRefGoogle ScholarPubMed
Hasegawa, M. (2006). Biochemistry and molecular biology of tauopathies. Neuropathology 26, 484–490.CrossRefGoogle ScholarPubMed
Hata, K., Fujitani, M., Yasuda, Y., et al. (2006). RGMa inhibition promotes axonal growth and recovery after spinal cord injury. J Cell Biol 173, 47–58.CrossRefGoogle ScholarPubMed
Hatano, Y., Miura, I., Horiuchi, T., et al. (1997). Cerebellar myeloblastoma formation in CD7-positive, neural cell adhesion molecule (CD56)-positive acute myelogenous leukemia (M1). Ann Hematol 75, 125–128.CrossRefGoogle Scholar
Hauser, S. L., Waubant, E., Arnold, D. L., et al. (2008). B-cell depletion with rituximab in relapsing-remitting multiple sclerosis. N Engl J Med. 358, 676–688.CrossRefGoogle ScholarPubMed
Hayes, G. M., Woodroofe, M. N., and Cuzner, M. L. (1987). Microglia are the major cell type expressing MHC class II in human white matter. J Neurol Sci 80, 25–37.CrossRefGoogle ScholarPubMed
Haynes, R. L., Billiards, S. S., Borenstein, N. S., Volpe, J. J. and Kinney, H. C. (2008). Diffuse axonal injury in periventricular leukomalacia as determined by apoptotic marker fractin. Pediatr Res 63, 656–661.CrossRefGoogle ScholarPubMed
Haynes, R. L., Borenstein, N. S., DeSilva, T. M., et al. (2005). Axonal development in the cerebral white matter of the human fetus and infant. J Comp Neurol 484, 156–167.CrossRefGoogle ScholarPubMed
He, X. L., Bazan, J. F., McDermott, G., et al. (2003). Structure of the Nogo receptor ectodomain: a recognition module implicated in myelin inhibition. Neuron 38, 177–185.CrossRefGoogle ScholarPubMed
Hedtjarn, M., Mallard, C., Arvidsson, P., and Hagberg, H. (2005). White matter injury in the immature brain: role of interleukin-18. Neurosci Lett 373, 16–20.CrossRefGoogle ScholarPubMed
Heese, K., Hock, C., and Otten, U. (1998). Inflammatory signals induce neurotrophin expression in human microglial cells. J Neurochem 70, 699–707.CrossRefGoogle ScholarPubMed
Heick, A. and Skriver, E. (2000). Chlamydia pneumoniae-associated ADEM. Eur J Neurol 7, 435–438.CrossRefGoogle ScholarPubMed
Heine, S., Ebnet, J., Maysami, S., and Stangel, M. (2006). Effects of interferon-beta on oligodendroglial cells. J Neuroimmunol 177, 173–180.CrossRefGoogle ScholarPubMed
Hemmer, B., Nessler, S., Zhou, D., Kieseier, B., and Hartung, H. P. (2006). Immunopathogenesis and immunotherapy of multiple sclerosis. Nat Clin Pract Neurol 2, 201–211.CrossRefGoogle ScholarPubMed
Henneke, M., Combes, P., Diekmann, S., et al. (2008). GJA12 mutations are a rare cause of Pelizaeus-Merzbacher-like disease. Neurology 70, 744–745.CrossRefGoogle ScholarPubMed
Henson, J. W. (1994). Regulation of the glial-specific JC virus early promoter by the transcription factor Sp1. J Biol Chem 269, 1046–1050.Google ScholarPubMed
Herman, M. A. and Jahr, C. E. (2007). Extracellular glutamate concentration in hippocampal slice. J Neurosci 27, 9736–9741.CrossRefGoogle ScholarPubMed
Heumann, R., Korsching, S., Bandtlow, C., and Thoenen, H. (1987). Changes of nerve growth factor synthesis in nonneuronal cells in response to sciatic nerve transection. J Cell Biol 104, 1623–1631.CrossRefGoogle ScholarPubMed
Hildebrand, C., Remahl, S., Persson, H., and Bjartmar, C. (1993). Myelinated nerve fibres in the CNS. Prog Neurobiol 40, 319–384.CrossRefGoogle ScholarPubMed
Hilliard, B., Wilmen, A., Seidel, C., Liu, T.-S. T., Göke, R., and Chen, Y. (2001). Roles of TNF-related apoptosis-inducing ligand in experimental autoimmune encephalomyelitis. J Immunol 166, 1314–1319.CrossRefGoogle ScholarPubMed
Hirayama, M., Yokochi, T., Shimokata, K., Iida, M., and Fujiki, N. (1986). Induction of human leukocyte antigen-A, B, C and -DR on cultured human oligodendrocytes and astrocytes by human gamma-interferon. Neurosci Lett 72, 369–374.CrossRefGoogle Scholar
Hisahara, S., Araki, T., Sugiyama, F., et al. (2000). Targeted expression of baculovirus p35 caspase inhibitor in oligodendrocytes protects mice against autoimmune-mediated demyelination. EMBO J 19, 341–348.CrossRefGoogle ScholarPubMed
Hobart, J., Kalkers, N., Barkhof, F., Uitdehaag, B., Polman, C., and Thompson, A. (2004). Outcome measures for multiple sclerosis clinical trials: relative measurement precision of the Expanded Disability Status Scale and Multiple Sclerosis Functional Composite. Mult Scler 10, 41–46.CrossRefGoogle ScholarPubMed
Hobart, J., Lamping, D., Fitzpatrick, R., Riazi, A., and Thompson, A. (2001). The Multiple Sclerosis Impact Scale (MSIS-29): a new patient-based outcome measure. Brain 124, 962–973.CrossRefGoogle ScholarPubMed
Hoek, K. S., Kidd, G. J., Carson, J. H., and Smith, R. (1998). hnRNP A2 selectively binds the cytoplasmic transport sequence of myelin basic protein mRNA. Biochemistry 37, 7021–7029.CrossRefGoogle ScholarPubMed
Hoftberger, R., Aboul-Enein, F., Brueck, W., et al. (2004). Expression of major histocompatibility complex class I molecules on the different cell types in multiple sclerosis lesions. Brain Pathol 14, 43–50.CrossRefGoogle Scholar
Hoke, A. (2006). Mechanisms of disease: what factors limit the success of peripheral nerve regeneration in humans? Nat Clin Pract Neurol 2, 448–454.CrossRefGoogle ScholarPubMed
Hoke, A., Redett, R., Hameed, H., et al. (2006). Schwann cells express motor and sensory phenotypes that regulate axon regeneration. J Neurosci 26, 9646–9655.CrossRefGoogle ScholarPubMed
Hollmann, M., Hartley, M., and Heinemann, S. (1991). Ca2+ permeability of KA-AMPA-gated glutamate receptor channels depends on subunit composition. Science 252, 851–853.CrossRefGoogle ScholarPubMed
Holmes, G. (1907). On the relation between loss of function and structural change in focal lesions of the central nervous system, with special reference to secondary degeneration. Brain 29, 514–523.CrossRefGoogle Scholar
Holz, A., Schaeren-Wiemers, N., Schaefer, C., Pott, U., Colello, R. J., and Schwab, M. E. (1996). Molecular and developmental characterization of novel cDNAs of the myelin-associated/oligodendrocytic basic protein. J Neurosci 16, 467–477.CrossRefGoogle ScholarPubMed
Honke, K., Hirahara, Y., Dupree, J., et al. (2002). Paranodal junction formation and spermatogenesis require sulfoglycolipids. Proc Natl Acad Sci USA 99, 4227–4232.CrossRefGoogle ScholarPubMed
Hooper, D. C., Spitsin, S., Kean, R. B., et al. (1998). Uric acid, a natural scavenger of peroxynitrite, in experimental allergic encephalomyelitis and multiple sclerosis. Proc Natl Acad Sci USA 95, 675–680.CrossRefGoogle ScholarPubMed
Horky, L. L., Galimi, F., Gage, F. H., and Horner, P. J. (2006). Fate of endogenous stem/progenitor cells following spinal cord injury. J Comp Neurol 498, 525–538.CrossRefGoogle ScholarPubMed
Horner, P. J., Power, A. E., Kempermann, G., et al. (2000). Proliferation and differentiation of progenitor cells throughout the intact adult rat spinal cord. J Neurosci 20, 2218–2228.CrossRefGoogle ScholarPubMed
Horner, P. J., Thallmair, M., and Gage, F. H. (2002). Defining the NG2-expressing cell of the adult CNS. J Neurocytol 31, 469–480.CrossRefGoogle ScholarPubMed
Hortega, P. D. R. (1928). Tercera aportacion al cenocimiento morfologico e interpretacion funcional de la oligodendroglia. Mem Real Soc Esp Hist Nat 14, 40–122.Google Scholar
Horssen, J., Schreibelt, G., Bo, L., et al. (2006). NAD(P)H:quinone oxidoreductase 1 expression in multiple sclerosis lesions. Free Radic Biol Med 41, 311–317.CrossRefGoogle ScholarPubMed
Hovelmeyer, N., Hao, Z., Kranidioti, K., et al. (2005). Apoptosis of oligodendrocytes via Fas and TNF-R1 is a key event in the induction of experimental autoimmune encephalomyelitis. J Immunol 175, 5875–5884.CrossRefGoogle ScholarPubMed
Howe, C. L., Bieber, A. J., Warrington, A. E., Pease, L. R., and Rodriguez, M. (2004). Antiapoptotic signaling by a remyelination-promoting human antimyelin antibody. Neurobiol Dis 15, 120–131.CrossRefGoogle ScholarPubMed
Hsieh, S.-T., Kidd, G. J., Crawford, T. O., et al. (1994). Regional modulation of neurofilament organization by myelination in normal axons. J Neurosci 14, 6392–6401.CrossRefGoogle ScholarPubMed
Hu, Q. D., Ang, B. T., Karsak, M., et al. (2003). F3/contactin acts as a functional ligand for Notch during oligodendrocyte maturation. Cell 115, 163–175.CrossRefGoogle ScholarPubMed
Hu, X., Hicks, C. W., He, W., et al. (2006). Bace1 modulates myelination in the central and peripheral nervous system. Nat Neurosci 9, 1520–1525.CrossRefGoogle ScholarPubMed
Huang, J. K., Phillips, G. R., Roth, A. D., et al. (2005). Glial membranes at the node of Ranvier prevent neurite outgrowth. Science 310, 1813–1817.CrossRefGoogle ScholarPubMed
Huangfu, D., Maehr, R., Guo, W., et al. (2008). Induction of pluripotent stem cells by defined factors is greatly improved by small-molecule compounds. Nat Biotechnol 26, 795–797.CrossRefGoogle ScholarPubMed
Huber, A. B., Weinmann, O., Brosamle, C., Oertle, T., and Schwab, M. E. (2002). Patterns of Nogo mRNA and protein expression in the developing and adult rat and after CNS lesions. J Neurosci 22, 3553–3567.CrossRefGoogle ScholarPubMed
Hudson, L. D. (2003). Pelizaeus–Merzbacher disease and spastic paraplegia type 2: two faces of myelin loss from mutations in the same gene. J Child Neurol 18, 616–624.CrossRefGoogle ScholarPubMed
Hudson, L. D., Garbern, J. Y., and Kamholz, J. A. (2004). Pelizaeus–Merzbacher disease. In Myelin Biology and Disorders, Lazzarini, R. A., ed. (Amsterdam: Elsevier Academic Press), pp. 867–885.Google Scholar
Huitinga, I., Rooijen, N., Groot, C. J., Uitdehaag, B. M., and Dijkstra, C. D. (1990). Suppression of experimental allergic encephalomyelitis in Lewis rats after elimination of macrophages. J Exp Med 172, 1025–1033.CrossRefGoogle ScholarPubMed
Hulsebosch, C. E. (2002). Recent advances in pathophysiology and treatment of spinal cord injury. Adv Physiol Educ 26, 238–255.CrossRefGoogle ScholarPubMed
Hulshof, S., Montagne, L., Groot, C. J., and Valk, P. (2002). Cellular localization and expression patterns of interleukin-10, interleukin-4, and their receptors in multiple sclerosis lesions. Glia 38, 24–35.CrossRefGoogle ScholarPubMed
Husain, J. and Juurlink, B. H. (1995). Oligodendroglial precursor cell susceptibility to hypoxia is related to poor ability to cope with reactive oxygen species. Brain Res 698, 86–94.CrossRefGoogle ScholarPubMed
Huseby, E. S., Liggitt, D., Brabb, T., Schnabel, B., Ohlen, C., and Goverman, J. (2001). A pathogenic role for myelin-specific CD8(+) T cells in a model for multiple sclerosis. J Exp Med 194, 669–676.CrossRefGoogle Scholar
Ide, C., Tohyama, K., Yokota, R., Nitatori, T., and Onodera, S. (1983). Schwann cell basal lamina and nerve regeneration. Brain Res 288, 61–75.CrossRefGoogle ScholarPubMed
Iglesias, A., Bauer, J., Litzenburger, T., Schubart, A., and Linington, C. (2001). T- and B-cell responses to myelin oligodendrocyte glycoprotein in experimental autoimmune encephalomyelitis and multiple sclerosis. Glia 36, 220–234.CrossRefGoogle Scholar
Ikonomidou, C., Bosch, F., Miksa, M., et al. (1999). Blockade of NMDA receptors and apoptotic neurodegeneration in the developing brain. Science 283, 70–74.CrossRefGoogle ScholarPubMed
Inder, T. E., Anderson, N. J., Spencer, C., Wells, S., and Volpe, J. J. (2003). White matter injury in the premature infant: a comparison between serial cranial sonographic and MR findings at term. AJNR Am J Neuroradiol 24, 805–809.Google Scholar
Inouye, H., Ganser, A. L., and Kirschner, D. A. (1985). Shiverer and normal peripheral myelin compared: basic protein localization, membrane interactions, and lipid composition. J Neurochem 45, 1911–1922.CrossRefGoogle ScholarPubMed
Irvine, K. A. and Blakemore, W. F. (2008). Remyelination protects axons from demyelination-associated axon degeneration. Brain 131, 1464–1477.CrossRefGoogle ScholarPubMed
Ishibashi, T., Dakin, K. A., Stevens, B., et al. (2006). Astrocytes promote myelination in response to electrical impulses. Neuron 49, 823–832.CrossRefGoogle ScholarPubMed
Ishibashi, T., Ding, L., Ikenaka, K., et al. (2004). Tetraspanin protein CD9 is a novel paranodal component regulating paranodal junctional formation. J Neurosci 24, 96–102.CrossRefGoogle ScholarPubMed
Itoh, T., Beesley, J., Itoh, A., et al. (2002). AMPA glutamate receptor-mediated calcium signaling is transiently enhanced during development of oligodendrocytes. J Neurochem 81, 390–402.CrossRefGoogle ScholarPubMed
Itoyama, Y., Ohnishi, A., Tateishi, J., Kuroiwa, Y., and Webster, H. D. (1985). Spinal cord multiple sclerosis lesions in Japanese patients: Schwann cell remyelination occurs in areas that lack glial fibrillary acidic protein (GFAP). Acta Neuropathol Berl 65, 217–223.CrossRefGoogle Scholar
Itoyama, Y., Sternberger, N. H., Webster, H. D., Quarles, R. H., Cohen, S. R., and Richardson, E. P.. (1980). Immunocytochemical observations on the distribution of myelin associated glycoprotein and myelin basic protein in multiple sclerosis lesions. Ann Neurol 7, 167–177.CrossRefGoogle ScholarPubMed
Itoyama, Y., Webster, H. D., Richardson, E. P., and Trapp, B. D. (1983). Schwann cell remyelination of demyelinated axons in spinal cord multiple sclerosis lesions. Ann Neurol 14, 339–346.CrossRefGoogle ScholarPubMed
Jack, C., Antel, J. P., Bruck, W., and Kuhlmann, T. (2007). Contrasting potential of nitric oxide and peroxynitrite to mediate oligodendrocyte injury in multiple sclerosis. Glia 55, 926–934.CrossRefGoogle ScholarPubMed
Jacobs, J. M., Cavanagh, J. B., and Chen, F. C. K. (1972). Spinal subarachnoid injection of colchicine in rats. J Neurol Sci 17, 461–480.CrossRefGoogle ScholarPubMed
Jacobsen, M., Schweer, D., Ziegler, A., et al. (2000). A point mutation in PTPRC is associated with the development of multiple sclerosis. Nat Genet 26, 495–499.CrossRefGoogle ScholarPubMed
Jacobson, M. (1978). Developmental Neurobiology (New York: Plenum.)CrossRefGoogle Scholar
Jahng, A. W., Maricic, I., Pedersen, B., et al. (2001). Activation of natural killer T cells potentiates or prevents experimental autoimmune encephalomyelitis. J Exp Med 194, 1789–1799.CrossRefGoogle ScholarPubMed
Jeffery, N. D. and Blakemore, W. F. (1997). Locomotor deficits induced by experimental spinal cord demyelination are abolished by spontaneous remyelination. Brain 120, 27–37.CrossRefGoogle ScholarPubMed
Jensen, A. M. and Chiu, S. Y. (1993). Expression of glutamate receptor genes in white matter: developing and adult rat optic nerve. J Neurosci 13, 1664–1675.CrossRefGoogle ScholarPubMed
Jessell, T. M. (2000). Neuronal specification in the spinal cord: inductive signals and transcriptional codes. Nature Rev Genet 1, 20–29.CrossRefGoogle ScholarPubMed
Jessen, K. R. and Mirsky, R. (2005a). The origin and development of glial cells in peripheral nerves. Nat Rev Neurosci 6, 671–682.CrossRefGoogle ScholarPubMed
Jessen, K. R. and Mirsky, R. (2005b). The Schwann cell lineage. In Neuroglia, Kettenmann, H. and Ransom, B. R., eds. (New York: Oxford University Press), pp. 85–100.Google Scholar
Ji, B., Case, L. C., Liu, K., et al. (2008). Assessment of functional recovery and axonal sprouting in oligodendrocyte-myelin glycoprotein (OMgp) null mice after spinal cord injury. Mol Cell Neurosci 39, 258–267.CrossRefGoogle ScholarPubMed
Johansson, C. B., Youssef, S., Koleckar, K., et al. (2008). Extensive fusion of haematopoietic cells with Purkinje neurons in response to chronic inflammation. Nat Cell Biol 10, 575–583.CrossRefGoogle ScholarPubMed
John, G. R., Shankar, S. L., Shafit-Zagardo, B., et al. (2002). Multiple sclerosis: re-expression of a developmental pathway that restricts oligodendrocyte maturation. Nat Med 8, 1115–1121.CrossRefGoogle ScholarPubMed
Johns, T. G. and Bernard, C. C. (1997). Binding of complement component Clq to myelin oligodendrocyte glycoprotein: a novel mechanism for regulating CNS inflammation. Mol Immunol 34, 33–38.CrossRefGoogle ScholarPubMed
Johns, T. G. and Bernard, C. C. (1999). The structure and function of myelin oligodendrocyte glycoprotein. J Neurochem 72, 1–9.CrossRefGoogle ScholarPubMed
Johnson, P. W., Abramow-Newerly, W., Seilheimer, B., et al. (1989). Recombinant myelin-associated glycoprotein confers neural adhesion and neurite outgrowth function. Neuron 3, 377–385.CrossRefGoogle ScholarPubMed
Joseph, N. M., Mukouyama, Y. S., Mosher, J. T., et al. (2004). Neural crest stem cells undergo multilineage differentiation in developing peripheral nerves to generate endoneurial fibroblasts in addition to Schwann cells. Development 131, 5599–5612.CrossRefGoogle ScholarPubMed
Jurewicz, A., Biddison, W. E., and Antel, J. P. (1998). MHC-Class I-restricted lysis of human oligodendrocytes by myelin basic protein peptide-specific CD8 T lymphocytes. J Immunol 160, 3056–3059.Google ScholarPubMed
Jurewicz, A., Matysiak, M., Andrzejak, S., and Selmaj, K. (2006). TRAIL-induced cell death of human adult oligodendrocytes is mediated by JNK pathway. Glia 15, 158–166.CrossRefGoogle Scholar
Jurewicz, A., Matysiak, M., Tybor, K., Kilianek, L., Raine, C. S., and Selmaj, K. (2005). Tumor necrosis factor-induced death of adult human oligodendrocytes is mediated by apoptosis inducing factor. Brain 128, 2675–2688.CrossRefGoogle ScholarPubMed
Jurewicz, A., Matysiak, M., Tybor, K., and Selmaj, K. (2003). TNF-induced death of adult human oligodendrocytes is mediated by c-jun NH(2)-terminal kinase-3. Brain 126, 1358–1370.CrossRefGoogle ScholarPubMed
Juurlink, B. J. H. (1997). Response of glial cells to ischemia: roles of reactive oxygen species and glutathione. Neurosci Biobehav Rev 21, 151–166.CrossRefGoogle ScholarPubMed
Juurlink, B. J. H., Thorburne, S. K., and Hertz, L. (1998). Peroxide-scavenging deficit underlies oligodendrocyte susceptibility to oxidative stress. Glia 22, 371–378.3.0.CO;2-6>CrossRefGoogle ScholarPubMed
Kachar, B., Behar, T., and Dubois-Dalcq, M. (1986). Cell shape and motility of oligodendrocytes cultured without neurons. Cell Tissue Res 244, 27–38.CrossRefGoogle ScholarPubMed
Kadi, L., Selvaraju, R., Lys, P., Proudfoot, A. E., Wells, T. N., and Boschert, U. (2006). Differential effects of chemokines on oligodendrocyte precursor proliferation and myelin formation in vitro. J Neuroimmunol 174, 133–146.CrossRefGoogle ScholarPubMed
Kagawa, T., Ikenaka, K., Inoue, Y., et al. (1994). Glial cell degeneration and hypomyelination caused by overexpression of myelin proteolipid protein gene. Neuron 13, 427–442.CrossRefGoogle ScholarPubMed
Kagawa, T., Mekada, E., Shishido, Y., and Ikenaka, K. (1997). Immune system-related CD9 is expressed in mouse central nervous system myelin at a very late stage of myelination. J Neurosci Res 50, 312–320.3.0.CO;2-9>CrossRefGoogle Scholar
Kaindl, A. M. and Ikonomidou, C. (2007). Glutamate antagonists are neurotoxins for the developing brain. Neurotox Res 11, 203–218.CrossRefGoogle ScholarPubMed
Kakulas, B. A. (1999). A review of the neuropathology of human spinal cord injury with emphasis on special features. J Spinal Cord Med 22, 119–124.CrossRefGoogle ScholarPubMed
Kamasawa, N., Sik, A., Morita, M., et al. (2005). Connexin-47 and connexin-32 in gap junctions of oligodendrocyte somata, myelin sheaths, paranodal loops and Schmidt–Lanterman incisures: implications for ionic homeostasis and potassium siphoning. Neuroscience 136, 65–86.CrossRefGoogle ScholarPubMed
Kanwar, J. R., Kanwar, R. K., and Krissansen, G. W. (2004). Simultaneous neuroprotection and blockade of inflammation reverses autoimmune encephalomyelitis. Brain 127, 1313–1331.CrossRefGoogle ScholarPubMed
Kaplan, M. R., Cho, M. H., Ullian, E. M., Isom, L. L., Levinson, S. R., and Barres, B. A. (2001). Differential control of clustering of the sodium channels Na(v)1.2 and Na(v)1.6 at developing CNS nodes of Ranvier. Neuron 30, 105–119.CrossRefGoogle ScholarPubMed
Kaplan, M. R., Meyer-Franke, A., Lambert, S., et al. (1997). Induction of sodium channel clustering by oligodendrocytes. Nature 386, 724–728.CrossRefGoogle ScholarPubMed
Kapsimali, M., Kloosterman, W. P., Bruijn, E., Rosa, F., Plasterk, R. H., and Wilson, S. W. (2007). MicroRNAs show a wide diversity of expression profiles in the developing and mature central nervous system. Genome Biol 8, R173.CrossRefGoogle ScholarPubMed
Karadottir, R. and Attwell, D. (2007). Neurotransmitter receptors in the life and death of oligodendrocytes. Neuroscience 145, 1426–1438.CrossRefGoogle ScholarPubMed
Karadottir, R., Cavelier, P., Bergersen, L. H., and Attwell, D. (2005). NMDA receptors are expressed in oligodendrocytes and activated in ischaemia. Nature 438, 1162–1166.CrossRefGoogle ScholarPubMed
Karadottir, R., Hamilton, N. B., Bakiri, Y., and Attwell, D. (2008). Spiking and nonspiking classes of oligodendrocyte precursor glia in CNS white matter. Nat Neurosci 11, 450–456.CrossRefGoogle ScholarPubMed
Karimi-Abdolrezaee, S., Eftekharpour, E., and Fehlings, M. G. (2004). Temporal and spatial patterns of Kv1.1 and Kv1.2 protein and gene expression in spinal cord white matter after acute and chronic spinal cord injury in rats: implications for axonal pathophysiology after neurotrauma. Eur J Neurosci 19, 577–589.CrossRefGoogle ScholarPubMed
Karimi-Abdolrezaee, S., Eftekharpour, E., Wang, J., Morshead, C. M., and Fehlings, M. G. (2006). Delayed transplantation of adult neural precursor cells promotes remyelination and functional neurological recovery after spinal cord injury. J Neurosci 26, 3377–3389.CrossRefGoogle ScholarPubMed
Karoutzou, G., Emrich, H. M., and Dietrich, D. E. (2008). The myelin-pathogenesis puzzle in schizophrenia: a literature review. Mol Psychiatry 13, 245–260.CrossRefGoogle ScholarPubMed
Kassis, I., Grigoriadis, N., Gowda-Kurkalli, B., et al. (2008). Neuroprotection and immunomodulation with mesenchymal stem cells in chronic experimental autoimmune encephalomyelitis. Arch Neurol 65, 753–761.CrossRefGoogle ScholarPubMed
Kassmann, C. M., Lappe-Siefke, C., Baes, M., et al. (2007). Axonal loss and neuroinflammation caused by peroxisome-deficient oligodendrocytes. Nat Genet 39, 969–976.CrossRefGoogle ScholarPubMed
Kato, T., Kakiuchi, C., and Iwamoto, K. (2007). Comprehensive gene expression analysis in bipolar disorder. Can J Psychiatry 52, 763–771.CrossRefGoogle ScholarPubMed
Keirstead, H. S., Hasan, S. J., Muir, G. D., and Steeves, J. D. (1992). Suppression of the onset of myelination extends the permissive period for the functional repair of embryonic spinal cord. Proc Natl Acad Sci USA 89, 11664–11668.CrossRefGoogle ScholarPubMed
Keirstead, H. S., Nistor, G., Bernal, G., et al. (2005). Human embryonic stem cell-derived oligodendrocyte progenitor cell transplants remyelinate and restore locomotion after spinal cord injury. J Neurosci 25, 4694–4705.CrossRefGoogle ScholarPubMed
Kelland, E. E. and Toms, N. J. (2001). Group I metabotropic glutamate receptors limit AMPA receptor-mediated oligodendrocyte progenitor cell death. Eur J Pharmacol 424, R3–R4.CrossRefGoogle Scholar
Kelm, S., Pelz, A., Schauer, R., et al. (1994). Sialoadhesin, myelin-associated glycoprotein and CD22 define a new family of sialic acid-dependent adhesion molecules of the immunoglobulin superfamily. Curr Biol 4, 965–972.CrossRefGoogle ScholarPubMed
Kerlero de Rosbo, N., Hoffman, M., Mendel, I., et al. (1997). Predominance of the autoimmune response to myelin oligodendrocyte glycoprotein in multiple sclerosis: reactivity to the extracellular domain of MOG is directed against three main regions. Eur J Immunol 27, 3059–3069.CrossRefGoogle ScholarPubMed
Kerlero de Rosbo, N., Milo, R., Lees, M. B., Burger, D., Bernard, C. C., and Ben Nun, A. (1993). Reactivity to myelin antigens in multiple sclerosis. Peripheral blood lymphocytes respond predominantly to myelin oligodendrocyte glycoprotein. J Clin Invest 92, 2602–2608.CrossRefGoogle ScholarPubMed
Kerschensteiner, M., Gallmeier, E., Behrens, L., et al. (1999). Activated human T cells, B cells, and monocytes produce brain-derived neurotrophic factor in vitro and in inflammatory brain lesions: a neuroprotective role of inflammation? J Exp Med 189, 865–870.CrossRefGoogle Scholar
Kerschensteiner, M., Meinl, E., and Hohlfeld, R. (2009). Neuro-immune crosstalk in CNS diseases. Neuroscience 158, 1122–1132.CrossRefGoogle ScholarPubMed
Kessaris, N., Fogarty, M., Iannarelli, P., Grist, M., Wegner, M., and Richardson, W. D. (2006). Competing waves of oligodendrocytes in the forebrain and postnatal elimination of an embryonic lineage. Nat Neurosci 9, 173–179.CrossRefGoogle ScholarPubMed
Kessaris, N., Pringle, N., and Richardson, W. D. (2008). Specification of CNS glia from neural stem cells in the embryonic neuroepithelium. Philos Trans R Soc Lond B Biol Sci 363, 71–85.CrossRefGoogle ScholarPubMed
Kettenmann, H. and Ransom, B. R. (2005). The concept of neuroglia: a historical perspective. In Neuroglia, Kettenmann, H. and Ransom, B. R., eds. (New York: Oxford University Press), pp. 1–16.Google Scholar
Kettenmann, H. and Verkhratsky, A. (2008). Neuroglia: the 150 years after. Trends Neurosci 31, 653–659.CrossRefGoogle Scholar
Kidd, G. J., Andrews, S. B., and Trapp, B. D. (1996). Axons regulate the distribution of Schwann cell microtubules. J Neurosci 16, 946–954.CrossRefGoogle ScholarPubMed
Kidd, G. J., Hauer, P. E., and Trapp, B. D. (1990). Axons modulate myelin protein messenger RNA levels during central nervous system myelination in vivo. J Neurosci Res 26, 409–418.CrossRefGoogle ScholarPubMed
Kidd, G. J., Yadav, V. K., Huang, P., et al. (2006). A dual tyrosine-leucine motif mediates myelin protein P0 targeting in MDCK cells. Glia 54, 135–145.CrossRefGoogle ScholarPubMed
Kierstead, H. S. and Blakemore, W. F. (1997). Identification of post-mitotic oligodendrocytes incapable of remyelination within the demyelinated adult spinal cord. J Neuropathol Exp Neurol 56, 1191–1201.CrossRefGoogle Scholar
Kim, H. J., DiBernardo, A. B., Sloane, J. A., et al. (2006). WAVE1 is required for oligodendrocyte morphogenesis and normal CNS myelination. J Neurosci 26, 5849–5859.CrossRefGoogle ScholarPubMed
Kim, J. E., Krichevsky, A., Grad, Y., et al. (2004a). Identification of many microRNAs that copurify with polyribosomes in mammalian neurons. Proc Natl Acad Sci USA 101, 360–365.CrossRefGoogle ScholarPubMed
Kim, J. E., Li, S., GrandPre, T., Qiu, D., and Strittmatter, S. M. (2003). Axon regeneration in young adult mice lacking Nogo-A/B. Neuron 38, 187–199.CrossRefGoogle ScholarPubMed
Kim, J. E., Liu, B. P., Park, J. H., and Strittmatter, S. M. (2004b). Nogo-66 receptor prevents raphespinal and rubrospinal axon regeneration and limits functional recovery from spinal cord injury. Neuron 44, 439–451.CrossRefGoogle ScholarPubMed
Kim, S. U., Moretto, G., and Shin, D. H. (1985). Expression of Ia antigens on the surface of human oligodendrocytes and astrocytes in culture. J Neuroimmunol 10, 141–149.CrossRefGoogle ScholarPubMed
Kim, Y. J., Park, H. J., Lee, G., et al. (2009). Neuroprotective effects of human mesenchymal stem cells on dopaminergic neurons through anti-inflammatory action. Glia 57, 13–23.CrossRefGoogle ScholarPubMed
Kimelberg, H. K., Goderie, S. K., Higman, S., Pang, S., and Waniewski, R. A. (1990). Swelling-induced release of glutamate, aspartate, and taurine from astrocyte cultures. J Neurosci 10, 1583–1591.CrossRefGoogle ScholarPubMed
Kimura, H., Weisz, A., Kurashima, Y., et al. (2000). Hypoxia response element of the human vascular endothelial growth factor gene mediates transcriptional regulation by nitric oxide: control of hypoxia-inducible factor-1 activity by nitric oxide. Blood 95, 189–197.Google ScholarPubMed
Kimura, M., Sato, M., Akatsuka, A., et al. (1989). Restoration of myelin formation by a single type of myelin basic protein in transgenic shiverer mice. Proc Natl Acad Sci USA 86, 5661–5665.CrossRefGoogle ScholarPubMed
Kinney, H. C. and Volpe, J. J. (2009). Perinatal panencephalopathy in premature infants: is it due to hypoxia-ischemia? In Brain Hypoxia and Ischemia, (Totowa, NJ: Humana Press).Google Scholar
Kinney, H. C., Brody, B. A., Kloman, A. S., and Gilles, F. H. (1988). Sequence of central nervous system myelination in human infancy. J Neuropathol Exp Neurol 47, 217–234.CrossRefGoogle ScholarPubMed
Kinney, H. C., Panigrahy, A., Newburger, J. W., Jonas, R. A., and Sleeper, L. A. (2005). Hypoxic-ischemic brain injury in infants with congenital heart disease dying after cardiac surgery. Acta Neuropathol 110, 563–578.CrossRefGoogle ScholarPubMed
Kirby, B. B., Takada, N., Latimer, A. J., et al. (2006). In vivo time-lapse imaging shows dynamic oligodendrocyte progenitor behavior during zebrafish development. Nat Neurosci 9, 1506–1511.CrossRefGoogle ScholarPubMed
Kirkpatrick, L. L., Witt, A. S., Payne, H. R., Shine, H. D., and Brady, S. T. (2001). Changes in microtubule stability and density in myelin-deficient shiverer mouse CNS axons. J Neurosci 21, 2288–2297.CrossRefGoogle ScholarPubMed
Kirschner, D. A. and Ganser, A. L. (1980). Compact myelin exists in the absence of basic protein in the shiverer mutant mouse. Nature 283, 207–210.CrossRefGoogle ScholarPubMed
Kitada, M. and Rowitch, D. H. (2006). Transcription factor co-expression patterns indicate heterogeneity of oligodendroglial subpopulations in adult spinal cord. Glia 54, 35–46.CrossRefGoogle ScholarPubMed
Kleopa, K. A., Orthmann, J. L., Enriquez, A., Paul, D. L., and Scherer, S. S. (2004). Unique distributions of the gap junction proteins connexin29, connexin32, and connexin47 in oligodendrocytes. Glia 47, 346–357.CrossRefGoogle ScholarPubMed
Klugmann, M., Schwab, M. H., Puhlhofer, A., et al. (1997). Assembly of CNS myelin in the absence of proteolipid protein. Neuron 18, 59–70.CrossRefGoogle ScholarPubMed
Koch, T., Brugger, T., Bach, A., Gennarini, G., and Trotter, J. (1997). Expression of the immunoglobulin superfamily cell adhesion molecule F3 by oligodendrocyte-lineage cells. Glia 19, 199–212.3.0.CO;2-V>CrossRefGoogle ScholarPubMed
Kocsis, J. D. and Waxman, S. G. (2007). Schwann cells and their precursors for repair of central nervous system myelin. Brain 130, 1978–1980.CrossRefGoogle ScholarPubMed
Kohama, I., Lankford, K. L., Preiningerova, J., White, F. A., Vollmer, T. L., and Kocsis, J. D. (2001). Transplantation of cryopreserved adult human Schwann cells enhances axonal conduction in demyelinated spinal cord. J Neurosci 21, 944–950.CrossRefGoogle ScholarPubMed
Kondo, T. and Raff, M. (2000). Oligodendrocyte precursor cells reprogrammed to become multipotential CNS stem cells [see comments]. Science 289, 1754–1757.CrossRefGoogle Scholar
Korbling, M. and Estrov, Z. (2003). Adult stem cells for tissue repair. N Engl J Med 349, 570–582.CrossRefGoogle ScholarPubMed
Kordeli, E., Davis, J. D., Trapp, B. D., and Bennett, V. (1990). An isoform of ankyrin is localized at nodes of Ranvier in myelinated axons of central and peripheral nerves. J Cell Biol 110, 1341–1352.CrossRefGoogle ScholarPubMed
Kordeli, E., Lambert, S., and Bennett, V. (1995). AnkyrinG. A new ankyrin gene with neural-specific isoforms localized at the axonal initial segment and node of Ranvier. J Biol Chem 270, 2352–2359.CrossRefGoogle ScholarPubMed
Kornek, B., Storch, M. K., Weissert, R., et al. (2000). Multiple sclerosis and chronic autoimmune encephalomyelitis: a comparative quantitative study of axonal injury in active, inactive, and remyelinated lesions. Am J Pathol 157, 267–276.CrossRefGoogle ScholarPubMed
Koticha, D., Maurel, P., Zanazzi, G., et al. (2006). Neurofascin interactions play a critical role in clustering sodium channels, ankyrin G and beta IV spectrin at peripheral nodes of Ranvier. Dev Biol 293, 1–12.CrossRefGoogle Scholar
Kotter, M. R., Li, W. W., Zhao, C., and Franklin, R. J. (2006). Myelin impairs CNS remyelination by inhibiting oligodendrocyte precursor cell differentiation. J Neurosci 26, 328–332.CrossRefGoogle ScholarPubMed
Kotter, M. R., Setzu, A., Sim, F. J., Rooijen, N., and Franklin, R. J. (2001). Macrophage depletion impairs oligodendrocyte remyelination following lysolecithin-induced demyelination. Glia 35, 204–212.CrossRefGoogle ScholarPubMed
Kottis, V., Thibault, P., Mikol, D., et al. (2002). Oligodendrocyte-myelin glycoprotein (OMgp) is an inhibitor of neurite outgrowth. J Neurochem 82, 1566–1569.CrossRefGoogle ScholarPubMed
Kramer, E. M., Schardt, A., and Nave, K. A. (2001). Membrane traffic in myelinating oligodendrocytes. Microsc Res Tech 52, 656–671.CrossRefGoogle ScholarPubMed
Krammer, P. H. (1999). CD95(APO-1/Fas)-mediated apoptosis: live and let die. Adv Immunol 71, 163–210.CrossRefGoogle ScholarPubMed
Krause, G., Winkler, L., Mueller, S. L., Haseloff, R. F., Piontek, J., and Blasig, I. E. (2008). Structure and function of claudins. Biochim Biophys Acta 1778, 631–645.CrossRefGoogle ScholarPubMed
Kress, G. J., Dineley, K. E., and Reynolds, I. J. (2002). The relationship between intracellular free iron and cell injury in cultured neurons, astrocytes, and oligodendrocytes. J Neurosci 22, 5848–5855.CrossRefGoogle ScholarPubMed
Krichevsky, A. M., King, K. S., Donahue, C. P., Khrapko, K., and Kosik, K. S. (2003). A microRNA array reveals extensive regulation of microRNAs during brain development. RNA 9, 1274–1281.CrossRefGoogle ScholarPubMed
Kroepfl, J. F. and Gardinier, M. V. (2001). Mutually exclusive apicobasolateral sorting of two oligodendroglial membrane proteins, proteolipid protein and myelin/oligodendrocyte glycoprotein, in Madin-Darby canine kidney cells. J Neurosci Res 66, 1140–1148.CrossRefGoogle ScholarPubMed
Kroepfl, J. F., Viise, L. R., Charron, A. J., Linington, C., and Gardinier, M. V. (1996). Investigation of myelin/oligodendrocyte glycoprotein membrane topology. J Neurochem 67, 2219–2222.CrossRefGoogle ScholarPubMed
Kuhlmann, T., Lucchinetti, C., Zettl, U. K., Bitsch, A., Lassmann, H., and Brück, W. (1999). Bcl-2-expressing oligodendrocytes in multiple sclerosis lesions. Glia 28, 34–39.3.0.CO;2-8>CrossRefGoogle ScholarPubMed
Kuhlmann, T., Miron, V., Cuo, Q., Wegner, C., Antel, J., and Bruck, W. (2008). Differentiation block of oligodendroglial progenitor cells as a cause for remyelination failure in chronic multiple sclerosis. Brain 131, 1749–1758.CrossRefGoogle ScholarPubMed
Kuhlmann, T., Wendling, U., Nolte, C., et al. (2002). Differential regulation of myelin phagocytosis by macrophages/microglia: involvement of target myelin, Fc receptors and activation by intravenous immunoglobulins. J Neurosci Res 67, 185–190.CrossRefGoogle ScholarPubMed
Kukley, M., Capetillo-Zarate, E., and Dietrich, D. (2007). Vesicular glutamate release from axons in white matter. Nat Neurosci 10, 311–320.CrossRefGoogle ScholarPubMed
Kukley, M., Kiladze, M., Tognatta, R., et al. (2008). Glial cells are born with synapses. FASEB J 22, 2957–2969.CrossRefGoogle ScholarPubMed
Kuner, T. and Schoepfer, R. (1996). Multiple structural elements determine subunit specificity of Mg2+ block in NMDA receptor channels. J Neurosci 16, 3549–3558.CrossRefGoogle ScholarPubMed
Kutzelnigg, A., Lucchinetti, C. F., Stadelmann, C., et al. (2005). Cortical demyelination and diffuse white matter injury in multiple sclerosis. Brain 128, 2705–2712.CrossRefGoogle ScholarPubMed
Kwiatkowski, D. J. (1999). Functions of gelsolin: motility, signaling, apoptosis, cancer. Curr Opin Cell Biol 11, 103–108.CrossRefGoogle ScholarPubMed
Kwiecien, J. M., O'Connor, L. T., Goetz, B. D., Delaney, K. H., Fletch, A. L., and Duncan, I. D. (1998). Morphological and morphometric studies of the dysmyelinating mutant, the Long Evans shaker rat. J Neurocytol 27, 581–591.CrossRefGoogle ScholarPubMed
Kwon, B. K., Oxland, T. R., and Tetzlaff, W. (2002). Animal models used in spinal cord regeneration research. Spine (Phila Pa 1976) 27, 1504–1510.CrossRefGoogle ScholarPubMed
Laabs, T., Carulli, D., Geller, H. M., and Fawcett, J. W. (2005). Chondroitin sulfate proteoglycans in neural development and regeneration. Curr Opin Neurobiol 15, 116–120.CrossRefGoogle Scholar
Labauge, P., Fogli, A., Niel, F., Rodriguez, D., and Boespflug-Tanguy, O. (2007). [CACH/VWM syndrome and leucodystrophies related to EIF2B mutations.] Rev Neurol (Paris) 163, 793–799.CrossRefGoogle ScholarPubMed
Lacas-Gervais, S., Guo, J., Strenzke, N., et al. (2004). BetaIVSigma1 spectrin stabilizes the nodes of Ranvier and axon initial segments. J Cell Biol 166, 983–990.CrossRefGoogle ScholarPubMed
Lachance, C., Arbour, N., Cashman, N. R., and Talbot, P. J. (1998). Involvement of aminopeptidase N (CD13) in infection of human neural cells by human coronavirus 229E. J Virol 72, 6511–6519.Google ScholarPubMed
Lagos-Quintana, M., Rauhut, R., Yalcin, A., Meyer, J., Lendeckel, W., and Tuschl, T. (2002). Identification of tissue-specific microRNAs from mouse. Curr Biol 12, 735–739.CrossRefGoogle ScholarPubMed
Lampasona, V., Franciotta, D., Furlan, R., et al. (2004). Similar low frequency of anti-MOG IgG and IgM in MS patients and healthy subjects. Neurology 62, 2092–2094.CrossRefGoogle ScholarPubMed
Langford, L. A., Porter, S., and Bunge, R. P. (1988). Immortalized rat Schwann cells produce tumours in vivo. J Neurocytol 17, 521–529.CrossRefGoogle ScholarPubMed
Lappe-Siefke, C., Goebbels, S., Gravel, M., et al. (2003). Disruption of Cnp1 uncouples oligodendroglial functions in axonal support and myelination. Nat Genet 33, 366–374.CrossRefGoogle ScholarPubMed
Lasiene, J., Matsui, A., Sawa, Y., Wong, F., and Horner, P. J. (2009). Age-related myelin dynamics revealed by increased oligodendrogenesis and short internodes. Aging Cell 8, 201–213.CrossRefGoogle ScholarPubMed
Lassmann, H. (1983). Comparative Neuropathology of Chronic Experimental Allergic Encephalomyelitis and Multiple Sclerosis (Berlin: Springer-Verlag).CrossRefGoogle ScholarPubMed
Lassmann, H. (2003). Axonal injury in multiple sclerosis. J Neurol Neurosurg Psychiatry 74, 695–697.CrossRefGoogle ScholarPubMed
Lassmann, H., Brück, W., Lucchinetti, C., and Rodriguez, M. (1997). Remyelination in multiple sclerosis. Mult Scler 3, 133–136.CrossRefGoogle ScholarPubMed
Lavdas, A., Franceschini, I., Dubois-Dalcq, M., and Matsas, R. (2006). Schwann cells genetically engineered to express PSA show enhanced migratory potential without impairment of their myelinating ability in vitro. Glia 53, 868–878.CrossRefGoogle ScholarPubMed
Lavdas, A. A., Papastefanaki, F., Thomaidou, D., and Matsas, R. (2008). Schwann cell transplantation for CNS repair. Curr Med Chem 15, 151–160.Google ScholarPubMed
Douarin, N. M. and Smith, J. (1988). Development of the peripheral nervous system from the neural crest. Annu Rev Cell Biol 4, 375–404.CrossRefGoogle ScholarPubMed
Lee, J., Gravel, M., Zhang, R., Thibault, P., and Braun, P. E. (2005). Process outgrowth in oligodendrocytes is mediated by CNP, a novel microtubule assembly myelin protein. J Cell Biol 170, 661–673.CrossRefGoogle ScholarPubMed
Lee, K. K., Repentigny, Y., Saulnier, R., Rippstein, P., Macklin, W. B., and Kothary, R. (2006). Dominant-negative beta1 integrin mice have region-specific myelin defects accompanied by alterations in MAPK activity. Glia 53, 836–844.CrossRefGoogle ScholarPubMed
Lee, R. C., Feinbaum, R. L., and Ambros, V. (1993). The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell 75, 843–854.CrossRefGoogle Scholar
Lee, S. C. and Raine, C. S. (1989). Multiple sclerosis: oligodendrocytes in active lesions do not express class II major histocompatibility complex molecules. J Neuroimmunol 25, 261–266.CrossRefGoogle Scholar
Lee, X., Yang, Z., Shao, Z., et al. (2007). NGF regulates the expression of axonal LINGO-1 to inhibit oligodendrocyte differentiation and myelination. J Neurosci 27, 220–225.CrossRefGoogle ScholarPubMed
Leegwater, P. A., Vermeulen, G., Konst, A. A., et al. (2001). Subunits of the translation initiation factor eIF2B are mutant in leukoencephalopathy with vanishing white matter. Nat Genet 29, 383–388.CrossRefGoogle ScholarPubMed
Lees, M. B. (1998). A history of proteolipids: a personal memoir. Neurochem Res 23, 261–271.CrossRefGoogle ScholarPubMed
Legrand, C., Ferraz, C., Clavel, M. C., and Rabie, A. (1986). Immunocytochemical localisation of gelsolin in oligodendroglia of the developing rabbit central nervous system. Brain Res 395, 231–235.CrossRefGoogle ScholarPubMed
Lemke, G. (1986). Molecular biology of the major myelin genes. Trends Neurosci 96, 266–270.CrossRefGoogle Scholar
Lena, J. Y., Legrand, C., Faivre-Sarrailh, C., Sarlieve, L. L., Ferraz, C., and Rabie, A. (1994). High gelsolin content of developing oligodendrocytes. Int J Dev Neurosci 12, 375–386.CrossRefGoogle ScholarPubMed
Lennon, V. A., Wingerchuk, D. M., Kryzer, T. J., et al. (2004). A serum autoantibody marker of neuromyelitis optica: distinction from multiple sclerosis. Lancet 364, 2106–2112.CrossRefGoogle ScholarPubMed
Leocani, L., Rovaris, M., Boneschi, F. M., et al. (2006). Multimodal evoked potentials to assess the evolution of multiple sclerosis: a longitudinal study. J Neurol Neurosurg Psychiatry 77, 1030–1035.CrossRefGoogle ScholarPubMed
Lepore, A. C. and Fischer, I. (2005). Lineage-restricted neural precursors survive, migrate, and differentiate following transplantation into the injured adult spinal cord. Exp Neurol 194, 230–242.CrossRefGoogle ScholarPubMed
Leuchtmann, E. A., Ratner, A. E., Vijitruth, R., Qu, Y., and McDonald, J. W. (2003). AMPA receptors are the major mediators of excitotoxic death in mature oligodendrocytes. Neurobiol Dis 14, 336–348.CrossRefGoogle ScholarPubMed
Levi, A. D. O. and Bunge, R. P. (1994). Studies of myelin formation after transplantation of human Schwann cells into the severe combined immunodeficient mouse. Exp Neurol 130, 41–52.CrossRefGoogle ScholarPubMed
Levine, J. M. and Reynolds, R. (1999). Activation and proliferation of endogenous oligodendrocyte precursor cells during ethidium bromide-induced demyelination. Exp Neurol 160, 333–347.CrossRefGoogle ScholarPubMed
Levine, J. M., Reynolds, R., and Fawcett, J. W. (2001). The oligodendrocyte precursor cell in health and disease. Trends Neurosci 24, 39–47.CrossRefGoogle ScholarPubMed
Levine, J. M., Stincone, F., and Lee, Y. S. (1993). Development and differentiation of glial precursor cells in the rat cerebellum. Glia 7, 307–321.CrossRefGoogle ScholarPubMed
Leviton, A. and Gilles, F. H. (1973). An epidemiologic study of perinatal telencephalic leucoencephalopathy in an autopsy population. J Neurol Sci 18, 53–66.CrossRefGoogle Scholar
Li, C., Tropak, M. B., Gerlai, R., et al. (1994). Myelination in the absence of myelin-associated glycoprotein. Nature 369, 747–750.CrossRefGoogle ScholarPubMed
Li, G. L., Brodin, G., Farooque, M., et al. (1996). Apoptosis and expression of Bcl-2 after compression trauma to rat spinal cord. J Neuropathol Exp Neurol 55, 280–289.CrossRefGoogle ScholarPubMed
Li, J., Baud, O., Vartanian, T., Volpe, J. J., and Rosenberg, P. A. (2005). Peroxynitrite generated by inducible nitric oxide synthetase and NADPH oxidase mediates microglia toxicity to oligodendrocytes. Proc Natl Acad Sci USA 102, 9936–9941.CrossRefGoogle ScholarPubMed
Li, J., Hertzberg, E. L., and Nagy, J. I. (1997). Connexin32 in oligodendrocytes and association with myelinated fibers in mouse and rat brain. J Comp Neurol 379, 571–591.3.0.CO;2-#>CrossRefGoogle ScholarPubMed
Li, M., Shibata, A., Li, C., et al. (1996). Myelin-associated glycoprotein inhibits neurite/axon growth and causes growth cone collapse. J Neurosci Res 46, 404–414.3.0.CO;2-K>CrossRefGoogle ScholarPubMed
Li, S., Kim, J. E., Budel, S., Hampton, T. G., and Strittmatter, S. M. (2005). Transgenic inhibition of Nogo-66 receptor function allows axonal sprouting and improved locomotion after spinal injury. Mol Cell Neurosci 29, 26–39.CrossRefGoogle ScholarPubMed
Li, S., Liu, B. P., Budel, S., et al. (2004). Blockade of Nogo-66, myelin-associated glycoprotein, and oligodendrocyte myelin glycoprotein by soluble Nogo-66 receptor promotes axonal sprouting and recovery after spinal injury. J Neurosci 24, 10511–10520.CrossRefGoogle ScholarPubMed
Li, S. and Strittmatter, S. M. (2003). Delayed systemic Nogo-66 receptor antagonist promotes recovery from spinal cord injury. J Neurosci 23, 4219–4227.CrossRefGoogle ScholarPubMed
Li, S., Mealing, G. A., Morley, P., and Stys, P. K. (1999). Novel injury mechanism in anoxia and trauma of spinal cord white matter: glutamate release via reverse Na+-dependent glutamate transport. J Neurosci 19, RC16.CrossRefGoogle ScholarPubMed
Li, W., Maeda, Y., Ming, X., et al. (2002). Apoptotic death following Fas activation in human oligodendrocyte hybrid cultures. J Neurosci Res 69, 189–196.CrossRefGoogle ScholarPubMed
Li, Y., Chen, J., Chen, X. G., et al. (2002). Human marrow stromal cell therapy for stroke in rat: neurotrophins and functional recovery. Neurology 59, 514–523.CrossRefGoogle ScholarPubMed
Li, Y., Chen, J., Wang, L., Lu, M., and Chopp, M. (2001a). Treatment of stroke in rat with intracarotid administration of marrow stromal cells. Neurology 56, 1666–1672.CrossRefGoogle ScholarPubMed
Li, Y., Chen, J., Wang, L., Zhang, L., Lu, M., and Chopp, M. (2001b). Intracerebral transplantation of bone marrow stromal cells in a 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine mouse model of Parkinson's disease. Neurosci Lett 316, 67–70.CrossRefGoogle Scholar
Li, Y., Chen, J., Zhang, C. L., et al. (2005). Gliosis and brain remodeling after treatment of stroke in rats with marrow stromal cells. Glia 49, 407–417.CrossRefGoogle ScholarPubMed
Li, Y., Field, P. M., and Raisman, G. (1997). Repair of adult rat corticospinal tract by transplants of olfactory ensheathing cells. Science 277, 2000–2002.CrossRefGoogle ScholarPubMed
Liebscher, T., Schnell, L., Schnell, D., et al. (2005). Nogo-A antibody improves regeneration and locomotion of spinal cord-injured rats. Ann Neurol 58, 706–719.CrossRefGoogle ScholarPubMed
Lin, S. C. and Bergles, D. E. (2004). Synaptic signaling between neurons and glia. Glia 47, 290–298.CrossRefGoogle ScholarPubMed
Lin, W., Bailey, S. L., Ho, H., et al. (2007). The integrated stress response prevents demyelination by protecting oligodendrocytes against immune-mediated damage. J Clin Invest 117, 448–456.CrossRefGoogle ScholarPubMed
Lin, W., Kemper, A., Dupree, J. L., Harding, H. P., Ron, D., and Popko, B. (2006). Interferon-gamma inhibits central nervous system remyelination through a process modulated by endoplasmic reticulum stress. Brain 129, 1306–1318.CrossRefGoogle ScholarPubMed
Linington, C., Bradl, M., Lassmann, H., Brunner, C., and Vass, K. (1988). Augmentation of demyelination in rat acute allergic encephalomyelitis by circulating mouse monoclonal antibodies directed against a myelin/oligodendrocyte glycoprotein. Am J Pathol 130, 443–454.Google ScholarPubMed
Linington, C., Engelhardt, B., Kapocs, G., and Lassmann, H. (1992). Induction of persistently demyelinated lesions in the rat following the repeated adoptive transfer of encephalitogenic T cells and demyelinating antibody. J Neuroimmunol 40, 219–224.CrossRefGoogle ScholarPubMed
Linington, C., Lassmann, H., Morgan, B. P., and Compston, D. A. S. (1989). Immunohistochemical localization of terminal complement component C9 in experimental allergic encephalomyelitis. Acta Neuropathol 79, 78–85.CrossRefGoogle Scholar
Linnington, C., Webb, M., and Woodhams, P. L. (1984). A novel myelin-associated glycoprotein defined by a mouse monoclonal antibody. J Neuroimmunol 6, 387–396.CrossRefGoogle ScholarPubMed
Lipton, S. A. (1986). Blockade of electrical-activity promotes the death of mammalian retinal ganglion-cells in culture. Proc Natl Acad Sci USA 83, 9774–9778.CrossRefGoogle ScholarPubMed
Lipton, S. A. and Rosenberg, P. A. (1994). Mechanisms of disease: excitatory amino acids as a final common pathway for neurologic disorders. N Engl J Med 330, 613–622.Google Scholar
Liu, B. P., Cafferty, W. B., Budel, S. O., and Strittmatter, S. M. (2006a). Extracellular regulators of axonal growth in the adult central nervous system. Philos Trans R Soc Lond B Biol Sci 361, 1593–1610.CrossRefGoogle ScholarPubMed
Liu, B. P., Fournier, A., GrandPre, T., and Strittmatter, S. M. (2002). Myelin-associated glycoprotein as a functional ligand for the Nogo-66 receptor. Science 297, 1190–1193.CrossRefGoogle ScholarPubMed
Liu, D., Liu, J., Sun, D., Alcock, N. W., and Wen, J. (2003). Spinal cord injury increases iron levels: catalytic production of hydroxyl radicals. Free Radic Biol Med 34, 64–71.CrossRefGoogle ScholarPubMed
Liu, D., Liu, J., and Wen, J. (1999). Elevation of hydrogen peroxide after spinal cord injury detected by using the Fenton reaction. Free Radic Biol Med 27, 478–482.CrossRefGoogle ScholarPubMed
Liu, J., Marino, M. W., Wong, G., et al. (1998). TNF is a potent anti-inflammatory cytokine in autoimmune-mediated demyelination. Nat Med 4, 78–83.CrossRefGoogle ScholarPubMed
Liu, J. S., Zhao, M. L., Brosnan, C. F., and Lee, S. C. (2001). Expression of inducible nitric oxide synthase and nitrotyrosine in multiple sclerosis lesions. Am J Pathol 158, 2057–2066.CrossRefGoogle ScholarPubMed
Liu, X. Z., Xu, X. M., Hu, R., et al. (1997). Neuronal and glial apoptosis after traumatic spinal cord injury. J Neurosci 17, 5395–5406.CrossRefGoogle ScholarPubMed
Liu, Y., Hao, W., Letiembre, M., et al. (2006b). Suppression of microglial inflammatory activity by myelin phagocytosis: role of p47-PHOX-mediated generation of reactive oxygen species. J Neurosci 26, 12904–12913.CrossRefGoogle ScholarPubMed
Lobsiger, C. S., Schweitzer, B., Taylor, V., and Suter, U. (2000). Platelet-derived growth factor-BB supports the survival of cultured rat Schwann cell precursors in synergy with neurotrophin-3. Glia 30, 290–300.3.0.CO;2-6>CrossRefGoogle ScholarPubMed
Loeliger, M., Watson, C. S., Reynolds, J. D., et al. (2003). Extracellular glutamate levels and neuropathology in cerebral white matter following repeated umbilical cord occlusion in the near term fetal sheep. Neuroscience 116, 705–714.CrossRefGoogle ScholarPubMed
Loevner, L. A., Shapiro, R. M., Grossman, R. I., Overhauser, J., and Kamholz, J. (1996). White matter changes associated with deletions of the long arm of chromosome 18 (18q- syndrome): a dysmyelinating disorder? AJNR Am J Neuroradiol 17, 1843–1848.Google Scholar
Lohse, D. C., Senter, H. J., Kauer, J. S., and Wohns, R. (1980). Spinal cord blood flow in experimental transient traumatic paraplegia. J Neurosurg 52, 335–345.CrossRefGoogle ScholarPubMed
Lopez-Munoz, F., Boya, J., and Alamo, C. (2006). Neuron theory, the cornerstone of neuroscience, on the centenary of the Nobel Prize award to Santiago Ramón y Cajal. Brain Res Bull 70, 391–405.CrossRefGoogle ScholarPubMed
LoPresti, P., Szuchet, S., Papasozomenos, S. C., Zinkowski, R. P., and Binder, L. I. (1995). Functional implications for the microtubule-associated protein tau: localization in oligodendrocytes. Proc Natl Acad Sci USA 92, 10369–10373.CrossRefGoogle ScholarPubMed
Love, S. (2006). Demyelinating diseases. J Clin Pathol 59, 1151–1159.CrossRefGoogle ScholarPubMed
Low, K., Culbertson, M., Bradke, F., Tessier-Lavigne, M., and Tuszynski, M. H. (2008). Netrin-1 is a novel myelin-associated inhibitor to axon growth. J Neurosci 28, 1099–1108.CrossRefGoogle ScholarPubMed
Low, S. H., Roche, P. A., Anderson, H. A., et al. (1998). Targeting of SNAP-23 and SNAP-25 in polarized epithelial cells. J Biol Chem 273, 3422–3430.CrossRefGoogle ScholarPubMed
Lu, Q. R., Sun, T., Zhu, Z., et al. (2002). Common developmental requirement for Olig function indicates a motor neuron/oligodendrocyte connection. Cell 109, 75–86.CrossRefGoogle ScholarPubMed
Lu, Q. R., Yuk, D., Alberta, J. A., et al. (2000). Sonic hedgehog – regulated oligodendrocyte lineage genes encoding bHLH proteins in the mammalian central nervous system. Neuron 25, 317–329.CrossRefGoogle ScholarPubMed
Lucchinetti, C., Brück, W., Parisi, J., Scheithauer, B., Rodriguez, M., and Lassmann, H. (1999). A quantitative analysis of oligodendrocytes in multiple sclerosis lesions. A study of 113 cases. Brain 122, 2279–2295.CrossRefGoogle ScholarPubMed
Lucchinetti, C., Brück, W., Parisi, J., Scheithauer, B., Rodriguez, M., and Lassmann, H. (2000). Heterogeneity of multiple sclerosis lesions: implications for the pathogenesis of demyelination. Ann Neurol 47, 707–717.3.0.CO;2-Q>CrossRefGoogle ScholarPubMed
Lucchinetti, C. F., Brück, W., Rodriguez, M., and Lassmann, H. (1996). Distinct patterns of multiple sclerosis pathology indicates heterogeneity in pathogenesis. Brain Pathol 6, 259–274.CrossRefGoogle ScholarPubMed
Lucchinetti, C. F., Mandler, R. N., McGavern, D., et al. (2002). A role for humoral mechanisms in the pathogenesis of Devic's neuromyelitis optica. Brain 125, 1450–1461.CrossRefGoogle ScholarPubMed
Luduena, R. F. (1998). Multiple forms of tubulin: different gene products and covalent modifications. Int Rev Cytol 178, 207–275.CrossRefGoogle ScholarPubMed
Ludwin, S. K. (1980). Chronic demyelination inhibits remyelination of the central nervous system. Lab Invest 43, 382–387.Google ScholarPubMed
Ludwin, S. K. (1988). Remyelination in the central nervous system and the peripheral nervous system. Adv Neurol 47, 215–254.Google ScholarPubMed
Ludwin, S. K. and Maitland, M. (1984). Long-term remyelination fails to reconstitute normal thickness of central myelin sheaths. J Neurol Sci 64, 193–198.CrossRefGoogle ScholarPubMed
Luetjens, C. M., Bui, N. T., Sengpiel, B., et al. (2000). Delayed mitochondrial dysfunction in excitotoxic neuron death: cytochrome c release and a secondary increase in superoxide production. J Neurosci 20, 5715–5723.CrossRefGoogle Scholar
Lumsden, C. E. (1951). Fundamental problems in the pathology of multiple sclerosis and allied demyelinating diseases. Br Med J1035–1043.CrossRef
Lunn, E. R., Scourfield, J., Keynes, R. J., and Stern, C. D. (1987). The neural tube origin of ventral root sheath cells in the chick embryo. Development 101, 247–254.Google ScholarPubMed
Lunn, K. F., Baas, P. W., and Duncan, I. D. (1997). Microtubule organization and stability in the oligodendrocyte. J Neurosci 17, 4921–4932.CrossRefGoogle ScholarPubMed
Luyt, K., Slade, T. P., Dorward, J. J., et al. (2007). Developing oligodendrocytes express functional GABA(B) receptors that stimulate cell proliferation and migration. J Neurochem 100, 822–840.CrossRefGoogle ScholarPubMed
Luyt, K., Varadi, A., Durant, C. F., and Molnar, E. (2006). Oligodendroglial metabotropic glutamate receptors are developmentally regulated and involved in the prevention of apoptosis. J Neurochem 99, 641–656.CrossRefGoogle Scholar
Luyt, K., Varadi, A., Halfpenny, C. A., Scolding, N. J., and Molnar, E. (2004). Metabotropic glutamate receptors are expressed in adult human glial progenitor cells. Biochem Biophys Res Commun 319, 120–129.CrossRefGoogle ScholarPubMed
Luzi, P., Zaka, M., Rao, H. Z., Curtis, M., Rafi, M. A., and Wenger, D. A. (2004). Generation of transgenic mice expressing insulin-like growth factor-1 under the control of the myelin basic protein promoter: increased myelination and potential for studies on the effects of increased IGF-1 on experimentally and genetically induced demyelination. Neurochem Res 29, 881–889.CrossRefGoogle ScholarPubMed
Macklin, W. B., Campagnoni, C. W., Deininger, P. L., and Gardinier, M. V. (1987a). Structure and expression of the mouse myelin proteolipid gene. J Neurosci Res 18, 383–394.CrossRefGoogle Scholar
Macklin, W. B., Gardinier, M. V., King, K. D., and Kampf, K. (1987b). An AG-GG transition at a splice site in the myelin PLP gene in jimpy mice results in the removal of an exon. FEBS Lett 223, 417–421.CrossRefGoogle Scholar
Madison, D. L., Krueger, W. H., Cheng, D., Trapp, B. D., and Pfeiffer, S. E. (1999). SNARE complex proteins, including the cognate pair VAMP-2 and syntaxin-4, are expressed in cultured oligodendrocytes. J Neurochem 72, 988–998.CrossRefGoogle ScholarPubMed
Madison, D. L., Krueger, W. H., Kim, T., and Pfeiffer, S. E. (1996). Differential expression of rab3 isoforms in oligodendrocytes and astrocytes. J Neurosci Res 45, 258–268.3.0.CO;2-C>CrossRefGoogle ScholarPubMed
Maeda, Y., Solanky, M., Menonna, J., Chapin, J., Li, W., and Dowling, P. (2001). Platelet-derived growth factor-alpha receptor-positive oligodendroglia are frequent in multiple sclerosis lesions. Ann Neurol 49, 776–785.CrossRefGoogle ScholarPubMed
Maggipinto, M., Rabiner, C., Kidd, G. J., Hawkins, A. J., Smith, R., and Barbarese, E. (2004). Increased expression of the MBP mRNA binding protein HnRNP A2 during oligodendrocyte differentiation. J Neurosci Res 75, 614–623.CrossRefGoogle ScholarPubMed
Magner, L. N. (2002). Problems in generation: preformation and epigenesis. In A History of the Life Sciences (Chicago, IL: CRC Press), pp. 152–204.Google Scholar
Maier, O., Heide, T., Johnson, R., Vries, H., Baron, W., and Hoekstra, D. (2006). The function of neurofascin155 in oligodendrocytes is regulated by metalloprotease-mediated cleavage and ectodomain shedding. Exp Cell Res 312, 500–511.CrossRefGoogle ScholarPubMed
Makeyev, E. V., Zhang, J., Carrasco, M. A., and Maniatis, T. (2007). The MicroRNA miR-124 promotes neuronal differentiation by triggering brain-specific alternative pre-mRNA splicing. Mol Cell 27, 435–448.CrossRefGoogle ScholarPubMed
Maletkovic, J., Schiffmann, R., Gorospe, J. R., et al. (2008). Genetic and clinical heterogeneity in eIF2B-related disorder. J Child Neurol 23, 205–215.CrossRefGoogle ScholarPubMed
Mallard, C., Welin, A. K., Peebles, D., Hagberg, H., and Kjellmer, I. (2003). White matter injury following systemic endotoxemia or asphyxia in the fetal sheep. Neurochem Res 28, 215–223.CrossRefGoogle ScholarPubMed
Mander, P., Borutaite, V., Moncada, S., and Brown, G. C. (2005). Nitric oxide from inflammatory-activated glia synergizes with hypoxia to induce neuronal death. J Neurosci Res 79, 208–215.CrossRefGoogle ScholarPubMed
Manitt, C., Wang, D., Kennedy, T. E., and Howland, D. R. (2006). Positioned to inhibit: netrin-1 and netrin receptor expression after spinal cord injury. J Neurosci Res 84, 1808–1820.CrossRefGoogle ScholarPubMed
Mann, S. A., Versmold, B., Marx, R., et al. (2008). Corticosteroids reverse cytokine-induced block of survival and differentiation of oligodendrocyte progenitor cells from rats. J Neuroinflammation 5, 39.CrossRefGoogle ScholarPubMed
Manning, S. M., Talos, D. M., Zhou, C., et al. (2008). NMDA receptor blockade with memantine attenuates white matter injury in a rat model of periventricular leukomalacia. J Neurosci 28, 6670–6678.CrossRefGoogle Scholar
Maragakis, N. J., Dietrich, J., Wong, V., et al. (2004). Glutamate transporter expression and function in human glial progenitors. Glia 45, 133–143.CrossRefGoogle ScholarPubMed
Marcus, J., Dupree, J. L., and Popko, B. (2000). Effects of galactolipid elimination on oligodendrocyte development and myelination. Glia 30, 319–328.3.0.CO;2-T>CrossRefGoogle ScholarPubMed
Markham, J. A. and Greenough, W. T. (2004). Experience-driven brain plasticity: beyond the synapse. Neuron Glia Biol 1, 351–363.CrossRefGoogle ScholarPubMed
Maro, G. S., Vermeren, M., Voiculescu, O., et al. (2004). Neural crest boundary cap cells constitute a source of neuronal and glial cells of the PNS. Nat Neurosci 7, 930–938.CrossRefGoogle ScholarPubMed
Marret, S., Mukendi, R., Gadisseux, J. F., Gressens, P., and Evrard, P. (1995). Effect of ibotenate on brain development: an excitotoxic mouse model of microgyria and post-hypoxic-like lesions. J Neuropathol Exp Neurol 54, 358–370.CrossRefGoogle ScholarPubMed
Marriott, M. P., Emery, B., Cate, H. S., et al. (2008). Leukemia inhibitory factor signaling modulates both central nervous system demyelination and myelin repair. Glia 56, 686–698.CrossRefGoogle ScholarPubMed
Martin, R., McFarland, H. F., and McFarlin, D. E. (1992). Immunological aspects of demyelinating diseases. Annu Rev Immunol 10, 153–187.CrossRefGoogle ScholarPubMed
Martini, R. (1994). Expression and functional roles of neural cell surface molecules and extracellular matrix components during development and regeneration of peripheral nerves. J Neurocytol 23, 1–28.CrossRefGoogle ScholarPubMed
Martini, R. (2005). Schwann cells and myelin. In Neuroglia, Ransom, B. R. and Kettenmann, H., eds. (Oxford: Oxford University Press), pp. 48–59.Google Scholar
Martini, R. and Schachner, M. (1986). Immunoelectron microscopic localization of neural cell adhesion molecules (L1, N-CAM, and MAG) and their shared carbohydrate epitope and myelin basic protein in developing sciatic nerve. J Cell Biol 103, 2439–2448.CrossRefGoogle ScholarPubMed
Martini, R. and Schachner, M. (1988). Immunoelectron microscopic localization of neural cell adhesion molecules (L1, N-CAM, and myelin-associated glycoprotein) in regenerating adult mouse sciatic nerve. J Cell Biol 106, 1735–1746.CrossRefGoogle Scholar
Martini, R. and Schachner, M. (1991). Complex expression pattern of tenascin during innervation of the posterior limb buds of the developing chicken. J Neurosci Res 28, 261–279.CrossRefGoogle ScholarPubMed
Martini, R., Fischer, S., Lopez-Vales, R., and David, S. (2008). Interactions between Schwann cells and macrophages in injury and inherited demyelinating disease. Glia 56, 1566–1577.CrossRefGoogle ScholarPubMed
Martini, R., Mohajeri, M. H., Kasper, S., Giese, K. P., and Schachner, M. (1995). Mice doubly deficient in the genes for P0 and myelin basic protein show that both proteins contribute to the formation of the major dense line in peripheral nerve myelin. J Neurosci 15, 4488–4495.CrossRefGoogle ScholarPubMed
Martini, R., Schachner, M., and Brushart, T. M. (1994). The L2/HNK-1 carbohydrate is preferentially expressed by previously motor axon-associated Schwann cells in reinnervated peripheral nerves. J Neurosci 14, 7180–7191.CrossRefGoogle ScholarPubMed
Martini, R., Schachner, M., and Faissner, A. (1990). Enhanced expression of the extracellular matrix molecule J1/tenascin in the regenerating adult mouse sciatic nerve. J Neurocytol 19, 601–616.CrossRefGoogle ScholarPubMed
Martini, R., Xin, Y., Schmitz, B., and Schachner, M. (1992). The L2/HNK-1 carbohydrate epitope is involved in the preferential outgrowth of motor neurons on ventral roots and motor nerves. Eur J Neurosci 4, 628–639.CrossRefGoogle ScholarPubMed
Martino, G. and Hartung, H. P. (1999). Immunopathogenesis of multiple sclerosis: the role of T cells. Curr Opin Neurol 12, 309–321.CrossRefGoogle ScholarPubMed
Mason, J. L., Jones, J. J., Taniike, M., Morell, P., Suzuki, K., and Matsushima, G. K. (2000). Mature oligodendrocyte apoptosis precedes IGF-1 production and oligodendrocyte progenitor accumulation and differentiation during demyelination/remyelination. J Neurosci Res 61, 251–262.3.0.CO;2-W>CrossRefGoogle ScholarPubMed
Mason, J. L., Suzuki, K., Chaplin, D. D., and Matsushima, G. K. (2001). Interleukin-1beta promotes repair of the CNS. J Neurosci 21, 7046–7052.CrossRefGoogle ScholarPubMed
Mason, J. L., Toews, A., Hostettler, J. D., et al. (2004). Oligodendrocytes and progenitors become progressively depleted within chronically demyelinated lesions. Am J Pathol 164, 1673–1682.CrossRefGoogle ScholarPubMed
Mata, M., Fink, D. J., Ernst, S. A., and Siegel, G. J. (1991). Immunocytochemical demonstration of Na+, K(+)-ATPase in internodal axolemma of myelinated fibers of rat sciatic and optic nerves. J Neurochem 57, 184–192.CrossRefGoogle ScholarPubMed
Mathey, E. K., Derfuss, T., Storch, M. K., et al. (2007). Neurofascin as a novel target for autoantibody-mediated axonal injury. J Exp Med 204, 2363–2372.CrossRefGoogle ScholarPubMed
Matloubian, M., Lo, C. G., Cinamon, G., et al. (2004). Lymphocyte egress from thymus and peripheral lymphoid organs is dependent on S1P receptor 1. Nature 427, 355–360.CrossRefGoogle ScholarPubMed
Matsumoto, Y., Kohyama, K., Aikawa, Y., et al. (1998). Role of natural killer cells and TCR gamma delta T cells in acute autoimmune encephalomyelitis. Eur J Immunol 28, 1681–1688.3.0.CO;2-T>CrossRefGoogle ScholarPubMed
Matute, C. (1998). Characteristics of acute and chronic kainate excitotoxic damage to the optic nerve. Proc Natl Acad Sci USA 95, 10229–10234.CrossRefGoogle ScholarPubMed
Matute, C., Alberdi, E., Domercq, M., et al. (2007). Excitotoxic damage to white matter. J Anat 210, 693–702.CrossRefGoogle ScholarPubMed
Matute, C., Sanchez-Gomez, M. V., Martinez-Millan, L., and Miledi, R. (1997). Glutamate receptor-mediated toxicity in optic nerve oligodendrocytes. Proc Natl Acad Sci USA 94, 8830–8835.CrossRefGoogle ScholarPubMed
Maurel, P., Einheber, S., Galinska, J., et al. (2007). Nectin-like proteins mediate axon Schwann cell interactions along the internode and are essential for myelination. J Cell Biol 178, 861–874.CrossRefGoogle ScholarPubMed
Maurer, M., Muller, M., Kobsar, I., Leonhard, C., Martini, R., and Kiefer, R. (2003). Origin of pathogenic macrophages and endoneurial fibroblast-like cells in an animal model of inherited neuropathy. Mol Cell Neurosci 23, 351–359.CrossRefGoogle Scholar
Mayer, M., Bhakoo, K., and Noble, M. (1994). Ciliary neurotrophic factor and leukemia inhibitory factor promote the generation, maturation and survival of oligodendrocytes in vitro. Development 120, 143–153.Google ScholarPubMed
Maysami, S., Nguyen, D., Zobel, F., Heine, S., Hopfner, M., and Stangel, M. (2006a). Oligodendrocyte precursor cells express a functional chemokine receptor CCR3: implications for myelination. J Neuroimmunol 178, 17–23.CrossRefGoogle ScholarPubMed
Maysami, S., Nguyen, D., Zobel, F., et al. (2006b). Modulation of rat oligodendrocyte precursor cells by the chemokine CXCL12. NeuroReport 17, 1187–1190.CrossRefGoogle ScholarPubMed
McCarran, W. J. and Goldberg, M. P. (2007). White matter axon vulnerability to AMPA/kainate receptor-mediated ischemic injury is developmentally regulated. J Neurosci 27, 4220–4229.CrossRefGoogle ScholarPubMed
McDonald, J. W., Althomsons, S. P., Hyrc, K. L., Choi, D. W., and Goldberg, M. P. (1998a). Oligodendrocytes from forebrain are highly vulnerable to AMPA/kainate receptor-mediated excitotoxicity. Nat Med 4, 291–297.CrossRefGoogle ScholarPubMed
McDonald, J. W., Levine, J. M., and Qu, Y. (1998b). Multiple classes of the oligodendrocyte lineage are highly vulnerable to excitotoxicity. NeuroReport 9, 2757–2762.CrossRefGoogle ScholarPubMed
McDonald, W. I. and Ohlrich, G. D. (1971). Quantitative anatomical measurements on single isolated fibres from the cat spinal cord. J Anat 110, 191–202.Google ScholarPubMed
McFarland, H. F. and Martin, R. (2007). Multiple sclerosis: a complicated picture of autoimmunity. Nat Immunol 8, 913–919.CrossRefGoogle ScholarPubMed
McGee, A. W., Yang, Y., Fischer, Q. S., Daw, N. W., and Strittmatter, S. M. (2005). Experience-driven plasticity of visual cortex limited by myelin and Nogo receptor. Science 309, 2222–2226.CrossRefGoogle ScholarPubMed
McKinnon, R. D., Piras, G., Ida, J. A.., and Dubois-Dalcq, M. (1993). A role for TGF-beta in oligodendrocyte differentiation. J Cell Biol 121, 1397–1407.CrossRefGoogle ScholarPubMed
McMorris, F. A. and Dubois-Dalcq, M. (1988). Insulin-like growth factor I promotes cell proliferation and oligodendroglial commitment in rat glial progenitor cells developing in vitro. J Neurosci Res 21, 199–209.CrossRefGoogle ScholarPubMed
McQuaid, S. and Cosby, S. L. (2002). An immunohistochemical study of the distribution of the measles virus receptors, CD46 and SLAM, in normal human tissues and subacute sclerosing panencephalitis. Lab Invest 82, 403–409.CrossRefGoogle ScholarPubMed
McTigue, D. M., Wei, P., and Stokes, B. T. (2001). Proliferation of NG2-positive cells and altered oligodendrocyte numbers in the contused rat spinal cord. J Neurosci 21, 3392–3400.CrossRefGoogle ScholarPubMed
Mead, R. J., Singhrao, S. K., Neal, J. W., Lassmann, H., and Morgan, B. P. (2002). The membrane attack complex of complement causes severe demyelination associated with acute axonal injury. J Immunol 168, 458–465.CrossRefGoogle ScholarPubMed
Medzhitov, R. and Janeway, C. A.. (1997). Innate immunity: impact on the adaptive immune response. Curr Opin Immunol 9, 4–9.CrossRefGoogle ScholarPubMed
Mei, L. and Xiong, W. C. (2008). Neuregulin 1 in neural development, synaptic plasticity and schizophrenia. Nat Rev Neurosci 9, 437–452.CrossRefGoogle Scholar
Meier, C., Parmantier, E., Brennan, A., Mirsky, R., and Jessen, K. R. (1999). Developing Schwann cells acquire the ability to survive without axons by establishing an autocrine circuit involving insulin-like growth factor, neurotrophin-3, and platelet-derived growth factor-BB. J Neurosci 19, 3847–3859.CrossRefGoogle ScholarPubMed
Mekki-Dauriac, S., Agius, E., Kan, P., and Cochard, P. (2002). Bone morphogenetic proteins negatively control oligodendrocyte precursor specification in the chick spinal cord. Development 129, 5117–5130.Google ScholarPubMed
Melendez-Vasquez, C., Carey, D. J., Zanazzi, G., Reizes, O., Maurel, P., and Salzer, J. L. (2005). Differential expression of proteoglycans at central and peripheral nodes of Ranvier. Glia 52, 301–308.CrossRefGoogle ScholarPubMed
Melendez-Vasquez, C. V., Rios, J. C., Zanazzi, G., Lambert, S., Bretscher, A., and Salzer, J. L. (2001). Nodes of Ranvier form in association with ezrin-radixin-moesin (ERM)-positive Schwann cell processes. Proc Natl Acad Sci USA 98, 1235–1240.CrossRefGoogle ScholarPubMed
Meletis, K., Barnabe-Heider, F., Carlen, M., et al. (2008). Spinal cord injury reveals multilineage differentiation of ependymal cells. PLoS Biol 6, e182.CrossRefGoogle ScholarPubMed
Menegoz, M., Gaspar, P., Bert, M., et al. (1997). Paranodin, a glycoprotein of neuronal paranodal membranes. Neuron 19, 319–331.CrossRefGoogle ScholarPubMed
Menichella, D. M., Goodenough, D. A., Sirkowski, E., Scherer, S. S., and Paul, D. L. (2003). Connexins are critical for normal myelination in the CNS. J Neurosci 23, 5963–5973.CrossRefGoogle ScholarPubMed
Menon, K. K., Piddlesden, S. J., and Bernard, C. C. A. (1997). Demyelinating antibodies to myelin oligodendrocyte glycoprotein and galactocerebroside induce degradation of myelin basic protein in isolated human myelin. J Neurochem 69, 214–222.CrossRefGoogle ScholarPubMed
Merkler, D., Boretius, S., Stadelmann, C., et al. (2005). Multicontrast MRI of remyelination in the central nervous system. NMR Biomed23, 7710–7718.Google ScholarPubMed
Merkler, D., Ernsting, T., Kerschensteiner, M., Bruck, W., and Stadelmann, C. (2006). A new focal EAE model of cortical demyelination: multiple sclerosis-like lesions with rapid resolution of inflammation and extensive remyelination. Brain 129, 1972–1983.CrossRefGoogle ScholarPubMed
Merkler, D., Metz, G. A., Raineteau, O., Dietz, V., Schwab, M. E., and Fouad, K. (2001). Locomotor recovery in spinal cord-injured rats treated with an antibody neutralizing the myelin-associated neurite growth inhibitor Nogo-A. J Neurosci 21, 3665–3673.CrossRefGoogle ScholarPubMed
Merrill, J. E. and Benveniste, E. N. (1996). Cytokines in inflammatory brain lesions: helpful and harmful. Trends Neurosci 19, 331–338.CrossRefGoogle ScholarPubMed
Merrill, J. E. and Zimmerman, R. P. (1991). Natural and induced cytotoxicity of oligodendrocytes by microglia is inhibitable by TGF beta. Glia 4, 327–331.CrossRefGoogle ScholarPubMed
Merrill, J. E., Ignarro, L. J., Sherman, M. P., Melinek, J., and Lane, T. E. (1993). Microglial cell cytotoxicity of oligodendrocytes is mediated through nitric oxide. J Immunol 151, 2132–2141.Google ScholarPubMed
Meucci, O., Fatatis, A., Holzwarth, J. A., and Miller, R. J. (1996). Developmental regulation of the toxin sensitivity of Ca(2+)-permeable AMPA receptors in cortical glia. J Neurosci 16, 519–530.CrossRefGoogle ScholarPubMed
MeyerFranke, A., Kaplan, M. R., Pfrieger, F. W., and Barres, B. A. (1995). Characterization of the signaling interactions that promote the survival and growth of developing retinal ganglion cells in culture. Neuron 15, 805–819.CrossRefGoogle Scholar
Mi, S., Hu, B., Hahm, K., et al. (2007). LINGO-1 antagonist promotes spinal cord remyelination and axonal integrity in MOG-induced experimental autoimmune encephalomyelitis. Nature Med 10, 1228–1233.CrossRefGoogle Scholar
Mi, S., Lee, X., Shao, Z., et al. (2004). LINGO-1 is a component of the Nogo-66 receptor/p75 signaling complex. Nat Neurosci 7, 221–228.CrossRefGoogle ScholarPubMed
Mi, S., Miller, R. H., Lee, X., et al. (2005). LINGO-1 negatively regulates myelination by oligodendrocytes. Nat Neurosci 8, 745–751.CrossRefGoogle ScholarPubMed
Micevych, P. E. and Abelson, L. (1991). Distribution of mRNAs coding for liver and heart gap junction proteins in the rat central nervous system. J Comp Neurol 305, 96–118.CrossRefGoogle ScholarPubMed
Michailov, G. V., Sereda, M. W., Brinkmann, B. G., et al. (2004). Axonal neuregulin-1 regulates myelin sheath thickness. Science 304, 700–703.CrossRefGoogle ScholarPubMed
Micu, I., Jiang, Q., Coderre, E., et al. (2006). NMDA receptors mediate calcium accumulation in myelin during chemical ischaemia. Nature 439, 988–992.Google ScholarPubMed
Miller, D. J. and Rodriguez, M. (1995). A monoclonal autoantibody that promotes central nervous system remyelination in a model of multiple sclerosis is a natural autoantibody encoded by germline immunoglobulin genes. J Immunol 154, 2460–2469.Google Scholar
Miller, R. H. (2002). Regulation of oligodendrocyte development in the vertebrate CNS. Prog Neurobiol 67, 451–467.CrossRefGoogle ScholarPubMed
Miller, R. H. (2005). Dorsally derived oligodendrocytes come of age. Neuron 45, 1–3.CrossRefGoogle ScholarPubMed
Miller, R. H., Dinsio, K. J., Wang, R. Z., Geertman, R., Maier, C. E., and Hall, A. K. (2004). Patterning of spinal cord oligodendrocyte development by dorsally derived BMP4. J Neurosci Res 76, 9–19.CrossRefGoogle ScholarPubMed
Miller, S. P., Cozzio, C. C., Goldstein, R. B., et al. (2003). Comparing the diagnosis of white matter injury in premature newborns with serial MR imaging and transfontanel ultrasonography findings. AJNR Am J Neuroradiol 24, 1661–1669.Google Scholar
Milner, R. (1997). Understanding the molecular basis of cell migration; implications for clinical therapy in multiple sclerosis. Clin Sci (Lond) 92, 113–122.CrossRefGoogle ScholarPubMed
Milner, R. and Ffrench-Constant, C. (1994). A developmental analysis of oligodendroglial integrins in primary cells: changes in alpha v-associated beta subunits during differentiation. Development 120, 3497–3506.Google ScholarPubMed
Mimeault, M. and Batra, S. K. (2008). Recent progress on tissue-resident adult stem cell biology and their therapeutic implications. Stem Cell Rev 4, 27–49.CrossRefGoogle ScholarPubMed
Minuk, J. and Braun, P. E. (1996). Differential intracellular sorting of the myelin-associated glycoprotein isoforms. J Neurosci Res 44, 411–420.3.0.CO;2-I>CrossRefGoogle ScholarPubMed
Mirsky, R., Winter, J., Abney, E. R., Pruss, R. M., Gavrilovic, J., and Raff, M. C. (1980). Myelin-specific proteins and glycolipids in rat Schwann cells and oligodendrocytes in culture. J Cell Biol 84, 483–494.CrossRefGoogle ScholarPubMed
Miska, E. A., Alvarez-Saavedra, E., Townsend, M., et al. (2004). Microarray analysis of microRNA expression in the developing mammalian brain. Genome Biol 5, R68.CrossRefGoogle ScholarPubMed
Mitani, A. and Tanaka, K. (2003). Functional changes of glial glutamate transporter GLT-1 during ischemia: an in vivo study in the hippocampal CA1 of normal mice and mutant mice lacking GLT-1. J Neurosci 23, 7176–7182.CrossRefGoogle Scholar
Mitrovic, B., Ignarro, L. J., Montestruque, S., Smoll, A., and Merrill, J. E. (1994). Nitric oxide as a potential pathological mechanism in demyelination: its differential effects on primary glial cells in vitro. Neuroscience 61, 575–585.CrossRefGoogle ScholarPubMed
Mitsunaga, Y., Ciric, B., Keulen, V., et al. (2002). Direct evidence that a human antibody derived from patient serum can promote myelin repair in a mouse model of chronic-progressive demyelinating disease. FASEB J 16, 1325–1327.CrossRefGoogle Scholar
Miura, M., Asou, H., Kobayashi, M., and Uyemura, K. (1992). Functional expression of a full-length cDNA coding for rat neural cell adhesion molecule L1 mediates homophilic intercellular adhesion and migration of cerebellar neurons. J Biol Chem 267, 10752–10758.Google ScholarPubMed
Miyamoto, E., Kakiuchi, S., and Kakimoto, Y. (1974). In vitro and in vivo phosphorylation of myelin basic protein by cerebral protein kinase. Nature 249, 150–151.CrossRefGoogle ScholarPubMed
Miyamoto, Y., Yamauchi, J., and Tanoue, A. (2008). Cdk5 phosphorylation of WAVE2 regulates oligodendrocyte precursor cell migration through nonreceptor tyrosine kinase Fyn. J Neurosci 28, 8326–8337.CrossRefGoogle ScholarPubMed
Mizuno, T., Sawada, M., Suzumura, A., and Marunouchi, T. (1994). Expression of cytokines during glial differentiation. Brain Res 656, 141–146.CrossRefGoogle ScholarPubMed
Moffett, J. R., Ross, B., Arun, P., Madhavarao, C. N., and Namboodiri, A. M. (2007). N-Acetylaspartate in the CNS: from neurodiagnostics to neurobiology. Prog Neurobiol 81, 89–131.CrossRefGoogle ScholarPubMed
Monje, M. L., Toda, H., and Palmer, T. D. (2003). Inflammatory blockade restores adult hippocampal neurogenesis. Science 302, 1760–1765.CrossRefGoogle ScholarPubMed
Montag, D., Giese, K. P., Bartsch, U., et al. (1994). Mice deficient for the myelin-associated glycoprotein show subtle abnormalities in myelin. Neuron 13, 229–246.CrossRefGoogle ScholarPubMed
Montague, P., McCallion, A. S., Davies, R. W., and Griffiths, I. R. (2006). Myelin-associated oligodendrocytic basic protein: a family of abundant CNS myelin proteins in search of a function. Dev Neurosci 28, 479–487.CrossRefGoogle ScholarPubMed
Moreau-Fauvarque, C., Kumanogoh, A., Camand, E., et al. (2003). The transmembrane semaphorin Sema4D/CD100, an inhibitor of axonal growth, is expressed on oligodendrocytes and upregulated after CNS lesion. J Neurosci 23, 9229–9239.CrossRefGoogle ScholarPubMed
Moretto, G., Xu, R. Y., and Kim, S. U. (1993). CD44 expression in human astrocytes and oligodendrocytes in culture. J Neuropathol Exp Neurol 52, 419–423.CrossRefGoogle ScholarPubMed
Morita, K., Sasaki, H., Fujimoto, K., Furuse, M., and Tsukita, S. (1999). Claudin-11/OSP-based tight junctions of myelin sheaths in brain and Sertoli cells in testis. J Cell Biol 145, 579–588.CrossRefGoogle ScholarPubMed
Moriya, T., Hassan, A. Z., Young, W., and Chesler, M. (1994). Dynamics of extracellular calcium activity following contusion of the rat spinal cord. J Neurotrauma 11, 255–263.CrossRefGoogle ScholarPubMed
Morris, J. K., Lin, W., Hauser, C., Marchuk, Y., Getman, D., and Lee, K. F. (1999). Rescue of the cardiac defect in ErbB2 mutant mice reveals essential roles of ErbB2 in peripheral nervous system development. Neuron 23, 273–283.CrossRefGoogle ScholarPubMed
Morrissey, T. K., Kleitman, N., and Bunge, R. P. (1995). Human Schwann cells in vitro. II. Myelination of sensory axons following extensive purification and heregulin-induced expansion. J Neurobiol 28, 190–201.CrossRefGoogle ScholarPubMed
Moscoso, L. M. and Sanes, J. R. (1995). Expression of four immunoglobulin superfamily adhesion molecules (L1, Nr-CAM/Bravo, neurofascin/ABGP, and N-CAM) in the developing mouse spinal cord. J Comp Neurol 352, 321–334.CrossRefGoogle ScholarPubMed
Moser, H. W., Mahmood, A., and Raymond, G. V. (2007). X-linked adrenoleukodystrophy. Nat Clin Pract Neurol 3, 140–151.CrossRefGoogle ScholarPubMed
Moser, H. W., Raymond, G. V., and Dubey, P. (2005). Adrenoleukodystrophy: new approaches to a neurodegenerative disease. J Am Med Assoc 294, 3131–3134.CrossRefGoogle ScholarPubMed
Mostov, K. E., Verges, M., and Altschuler, Y. (2000). Membrane traffic in polarized epithelial cells. Curr Opin Cell Biol 12, 483–490.CrossRefGoogle ScholarPubMed
Mronga, T., Stahnke, T., Goldbaum, O., and Richter-Landsberg, C. (2004). Mitochondrial pathway is involved in hydrogen-peroxide-induced apoptotic cell death of oligodendrocytes. Glia 46, 446–455.CrossRefGoogle ScholarPubMed
Mueller, M., Leonhard, C., Wacker, K., et al. (2003). Macrophage response to peripheral nerve injury: the quantitative contribution of resident and hematogenous macrophages. Lab Invest 83, 175–185.CrossRefGoogle ScholarPubMed
Mukhopadhyay, G., Doherty, P., Walsh, F. S., Crocker, P. R., and Filbin, M. T. (1994). A novel role for myelin-associated glycoprotein as an inhibitor of axonal regeneration. Neuron 13, 757–767.CrossRefGoogle ScholarPubMed
Munro, T. P., Magee, R. J., Kidd, G. J., et al. (1999). Mutational analysis of a heterogeneous nuclear ribonucleoprotein A2 response element for RNA trafficking. J Biol Chem 274, 34389–34395.CrossRefGoogle ScholarPubMed
Murru, M. R., Vannelli, A., Marrosu, G., et al. (2006). A novel Cx32 mutation causes X-linked Charcot-Marie-Tooth disease with brainstem involvement and brain magnetic resonance spectroscopy abnormalities. Neurol Sci 27, 18–23.CrossRefGoogle ScholarPubMed
Nagy, J. I., Ionescu, A. V., Lynn, B. D., and Rash, J. E. (2003). Connexin29 and connexin32 at oligodendrocyte and astrocyte gap junctions and in myelin of the mouse central nervous system. J Comp Neurol 464, 356–370.CrossRefGoogle ScholarPubMed
Nagy, Z., Westerberg, H., and Klingberg, T. (2004). Maturation of white matter is associated with the development of cognitive functions during childhood. J Cogn Neurosci 16, 1227–1233.CrossRefGoogle ScholarPubMed
Nait-Oumesmar, B., Picard-Riera, N., Kerninon, C., et al. (2007). Activation of the subventricular zone in multiple sclerosis: evidence for early glial progenitors. Proc Natl Acad Sci USA 104, 4694–4699.CrossRefGoogle ScholarPubMed
Nakagawa, M., Koyanagi, M., Tanabe, K., et al. (2008). Generation of induced pluripotent stem cells without Myc from mouse and human fibroblasts. Nat Biotechnol 26, 101–106.CrossRefGoogle ScholarPubMed
Nakahara, J., Seiwa, C., Tan-Takeuchi, K., et al. (2005). Involvement of CD45 in central nervous system myelination. Neurosci Lett. 379, 116–121.CrossRefGoogle ScholarPubMed
Nakajima, K. and Kohsaka, S. (2001). Microglia: activation and their significance in the central nervous system. J Biochem 130, 169–175.CrossRefGoogle ScholarPubMed
Namboodiri, A. M., Peethambaran, A., Mathew, R., et al. (2006). Canavan disease and the role of N-acetylaspartate in myelin synthesis. Mol Cell Endocrinol 252, 216–223.CrossRefGoogle ScholarPubMed
Narayanan, S. P., Flores, A. I., Wang, F., and Macklin, W. B. (2009). Akt signals through the mammalian target of rapamycin pathway to regulate CNS myelination. J Neurosci 29, 6860–6870.CrossRefGoogle ScholarPubMed
Naruse, M., Nakahira, E., Miyata, T., Hitoshi, S., Ikenaka, K., and Bansal, R. (2006). Induction of oligodendrocyte progenitors in dorsal forebrain by intraventricular microinjection of FGF-2. Dev Biol 297, 262–273.CrossRefGoogle ScholarPubMed
Nataf, S., Carroll, S. L., Wetsel, R. A., Szalai, A. J., and Barnum, S. R. (2000). Attenuation of experimental autoimmune demyelination in complement-deficient mice. J Immunol 165, 5867–5873.CrossRefGoogle ScholarPubMed
Nave, K. A. (2008). Neuroscience: an ageing view of myelin repair. Nature 455, 478–479.CrossRefGoogle ScholarPubMed
Nave, K. A. and Griffiths, I. (2004). Models of Pelizaeus-Merzbacher disease. In Myelin Biology and Disorders, Lazzarini, R. A., ed. (Amsterdam: Elsevier Academic Press), pp. 1125–1142.Google Scholar
Nave, K. A. and Salzer, J. L. (2006). Axonal regulation of myelination by neuregulin 1. Curr Opin Neurobiol 16, 492–500.CrossRefGoogle ScholarPubMed
Nave, K. A. and Trapp, B. D. (2008). Axon-glial signaling and the glial support of axon function. Annu Rev Neurosci 31, 535–561.CrossRefGoogle ScholarPubMed
Nave, K. A., Lai, C., Bloom, F. E., and Milner, R. J. (1987). Splice site selection in the proteolipid protein (PLP) gene transcript and primary structure of the DM-20 protein of central nervous system myelin. Proc Natl Acad Sci USA 84, 5665–5669.CrossRefGoogle ScholarPubMed
Nave, K. A., Sereda, M. W., and Ehrenreich, H. (2007). Mechanisms of disease: inherited demyelinating neuropathies – from basic to clinical research. Nat Clin Pract Neurol 3, 453–464.CrossRefGoogle ScholarPubMed
Nedergaard, M., Takano, T., and Hansen, A. J. (2002). Beyond the role of glutamate as a neurotransmitter. Nat Rev Neurosci 3, 748–755.CrossRefGoogle ScholarPubMed
Neuhaus, O., Kieseier, B. C., and Hartung, H. P. (2006). Mitoxantrone in multiple sclerosis. Adv Neurol 98, 293–302.Google ScholarPubMed
Neumann, B., Machleidt, T., Lifka, A., et al. (1996). Crucial role of 55-kilodalton TNF receptor in TNF-induced adhesion molecule expression and leukocyte organ infiltration. J Immunol 156, 1587–1593.Google ScholarPubMed
Nicolay, D. J., Doucette, J. R., and Nazarali, A. J. (2007). Transcriptional control of oligodendrogenesis. Glia 55, 1287–1299.CrossRefGoogle ScholarPubMed
Niederlander, C. and Lumsden, A. (1996). Late emigrating neural crest cells migrate specifically to the exit points of cranial branchiomotor nerves. Development 122, 2367–2374.Google ScholarPubMed
Niederost, B., Oertle, T., Fritsche, J., McKinney, R. A., and Bandtlow, C. E. (2002). Nogo-A and myelin-associated glycoprotein mediate neurite growth inhibition by antagonistic regulation of RhoA and Rac1. J Neurosci 22, 10368–10376.CrossRefGoogle ScholarPubMed
Niehaus, A., Shi, J., Grzenkowski, M., et al. (2000). Patients with active relapsing-remitting multiple sclerosis synthesize antibodies recognizing oligodendrocyte progenitor cell surface protein: implications for remyelination. Ann Neurol 48, 362–371.3.0.CO;2-6>CrossRefGoogle ScholarPubMed
Nielsen, J. A., Maric, D., Lau, P., Barker, J. L., and Hudson, L. D. (2006). Identification of a novel oligodendrocyte cell adhesion protein using gene expression profiling. J Neurosci 26, 9881–9891.CrossRefGoogle ScholarPubMed
Niemann, A., Berger, P., and Suter, U. (2006). Pathomechanisms of mutant proteins in Charcot-Marie-Tooth disease. Neuromolecular Med 8, 217–242.CrossRefGoogle ScholarPubMed
Nikizad, H., Yon, J. H., Carter, L. B., and Jevtovic-Todorovic, V. (2007). Early exposure to general anesthesia causes significant neuronal deletion in the developing rat brain. Ann N Y Acad Sci 1122, 69–82.CrossRefGoogle ScholarPubMed
Nimmerjahn, A., Kirchhoff, F., and Helmchen, F. (2005). Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo. Science 308, 1314–1318.CrossRefGoogle ScholarPubMed
Nishiyama, A. (2007). Polydendrocytes: NG2 cells with many roles in development and repair of the CNS. Neuroscientist 13, 62–76.CrossRefGoogle ScholarPubMed
Nishiyama, A., Komitova, M., Suzuki, R., and Zhu, X. (2009). Polydendrocytes (NG2 cells): multifunctional cells with lineage plasticity. Nat Rev Neurosci 10, 9–22.CrossRefGoogle ScholarPubMed
Nishiyama, A., Watanabe, M., Yang, Z., and Bu, J. (2002). Identity, distribution, and development of polydendrocytes: NG2-expressing glial cells. J Neurocytol 31, 437–455.CrossRefGoogle ScholarPubMed
Noakes, P. G. and Bennett, M. R. (1987). Growth of axons into developing muscles of the chick forelimb is preceded by cells that stain with Schwann cell antibodies. J Comp Neurol 259, 330–347.CrossRefGoogle ScholarPubMed
Noble, L. J. and Wrathall, J. R. (1989). Correlative analyses of lesion development and functional status after graded spinal cord contusive injuries in the rat. Exp Neurol 103, 34–40.CrossRefGoogle ScholarPubMed
Nodari, A., Previtali, S. C., Dati, G., et al. (2008). Alpha6beta4 integrin and dystroglycan cooperate to stabilize the myelin sheath. J Neurosci 28, 6714–6719.CrossRefGoogle ScholarPubMed
Nodari, A., Zambroni, D., Quattrini, A., et al. (2007). Beta1 integrin activates Rac1 in Schwann cells to generate radial lamellae during axonal sorting and myelination. J Cell Biol 177, 1063–1075.CrossRefGoogle ScholarPubMed
Norenberg, M. D., Smith, J., and Marcillo, A. (2004). The pathology of human spinal cord injury: defining the problems. J Neurotrauma 21, 429–440.CrossRefGoogle ScholarPubMed
Northington, F. J., Traystman, R. J., Koehler, R. C., and Martin, L. J. (1999). GLT1, glial glutamate transporter, is transiently expressed in neurons and develops astrocyte specificity only after midgestation in the ovine fetal brain. J Neurobiol 39, 515–526.3.0.CO;2-U>CrossRefGoogle ScholarPubMed
Norton, W. T. (1984). Oligodendroglia. (New York: Plenum Press.)CrossRefGoogle Scholar
Norton, W. T. (1996). Do oligodendrocytes divide? Neurochem Res 21, 495–503.CrossRefGoogle ScholarPubMed
Norton, W. T. and Cammer, W. (1984). Isolation and characterization of myelin. In Myelin, Morell, P., ed. (New York: Plenum Press), pp. 146–196.Google ScholarPubMed
Novgorodov, A. S., El-Alwani, M., Bielawski, J., Obeid, L. M., and Gudz, T. I. (2007). Activation of sphingosine-1-phosphate receptor S1P5 inhibits oligodendrocyte progenitor migration. FASEB J 21, 1503–1514.CrossRefGoogle ScholarPubMed
Nunes, M. C., Roy, N. S., Keyoung, H. M., et al. (2003). Identification and isolation of multipotential neural progenitor cells from the subcortical white matter of the adult human brain. Nat Med 9, 439–447.CrossRefGoogle ScholarPubMed
Nygren, J. M., Liuba, K., Breitbach, M., et al. (2008). Myeloid and lymphoid contribution to non-haematopoietic lineages through irradiation-induced heterotypic cell fusion. Nat Cell Biol 10, 584–592.CrossRefGoogle ScholarPubMed
Occhi, S., Zambroni, D., Del Carro, U., et al. (2005). Both laminin and Schwann cell dystroglycan are necessary for proper clustering of sodium channels at nodes of Ranvier. J Neurosci 25, 9418–9427.CrossRefGoogle ScholarPubMed
Odermatt, B., Wellershaus, K., Wallraff, A., et al. (2003). Connexin 47 (Cx47)-deficient mice with enhanced green fluorescent protein reporter gene reveal predominant oligodendrocytic expression of Cx47 and display vacuolized myelin in the CNS. J Neurosci 23, 4549–4559.CrossRefGoogle ScholarPubMed
Oertle, T., Haar, M. E., Bandtlow, C. E., et al. (2003). Nogo-A inhibits neurite outgrowth and cell spreading with three discrete regions. J Neurosci 23, 5393–5406.CrossRefGoogle ScholarPubMed
Ogata, J. and Feigin, I. (1975). Schwann cells and regenerated peripheral myelin in multiple sclerosis: an ultrastructural study. Neurology 25, 713–716.CrossRefGoogle Scholar
Oka, A., Belliveau, M. J., Rosenberg, P. A., and Volpe, J. J. (1993). Vulnerability of oligodendroglia to glutamate: pharmacology, mechanisms, and prevention. J Neurosci 13, 1441–1453.CrossRefGoogle Scholar
Oka, S., Honmou, O., Akiyama, Y., et al. (2004). Autologous transplantation of expanded neural precursor cells into the demyelinated monkey spinal cord. Brain Res 1030, 94–102.CrossRefGoogle ScholarPubMed
Oleszak, E. L., Zaczynska, E., Bhattacharjee, C., Butunoi, C., Ledigo, A., and Katsetos, C. (1998). Inducible nitric oxide synthase and nitrotyrosine are found in monocytes/macrophages and/or astrocytes in acute, but not in chronic multiple sclerosis. Clin Diagn Lab Immunol 5, 438–445.Google ScholarPubMed
Olney, J. W., Labruyere, J., Wang, G., Wozniak, D. F., Price, M. T., and Sesma, M. A. (1991). NMDA antagonist neurotoxicity: mechanism and prevention. Science 254, 1515–1518.CrossRefGoogle Scholar
Olson, J. K. and Miller, S. D. (2004). Microglia initiate central nervous system innate and adaptive immune responses through multiple TLRs. J Immunol 173, 3916–3924.CrossRefGoogle ScholarPubMed
Omari, K. M., John, G., Lango, R., and Raine, C. S. (2006). Role for CXCR2 and CXCL1 on glia in multiple sclerosis. Glia 53, 24–31.CrossRefGoogle ScholarPubMed
Omari, K. M., John, G. R., Sealfon, S. C., and Raine, C. S. (2005). CXC chemokine receptors on human oligodendrocytes: implications for multiple sclerosis. Brain 128, 1003–1015.CrossRefGoogle ScholarPubMed
Omlin, F. X., Webster, H. D., Palkovits, C. G., and Cohen, S. R. (1982). Immunocytochemical localization of basic protein in major dense line regions of central and peripheral myelin. J Cell Biol 95, 242–248.CrossRefGoogle ScholarPubMed
Ono, K., Bansal, R., Payne, J., Rutishauser, U., and Miller, R. H. (1995). Early development and dispersal of oligodendrocyte precursors in the embyonic chick spinal cord. Development 121, 1743–1754.Google Scholar
Ono, K., Fujisawa, H., Hirano, S., Norita, M., Tsumori, T., and Yasui, Y. (1997). Early development of the oligodendrocyte in the embryonic chick metencephalon. J Neurosci Res 48, 1–14.3.0.CO;2-I>CrossRefGoogle ScholarPubMed
Oosten, B. W., Barkhof, F., Truyen, L., Boringa, J. B., Bertelsmann, F. W., Blomberg, B. M., Woody, J. N., Hartung, H. P., and Polman, C. Het al. (1996). Increased MRI activity and immune activation in two multiple sclerosis patients treated with the monoclonal anti-tumor necrosis factor antibody cA2. Neurology 47, 1531–1534.CrossRefGoogle ScholarPubMed
Openshaw, H., Lund, B. T., Kashyap, A., et al. (2000). Peripheral blood stem cell transplantation in multiple sclerosis with busulfan and cyclophosphamide conditioning: report of toxicity and immunological monitoring. Biol Blood Marrow Transplant 6, 563–575.CrossRefGoogle ScholarPubMed
Orentas, D. M. and Miller, R. H. (1996). The origin of spinal cord oligodendrocytes is dependent on local influences from the notochord. Developmental Biology 177, 43–53.CrossRefGoogle ScholarPubMed
Orentas, D. M., Hayes, J. E., Dyer, K. L., and Miller, R. H. (1999). Sonic hedgehog signaling is required during the appearance of spinal cord oligodendrocyte precursors. Development 126, 2419–2429.Google ScholarPubMed
Orthmann-Murphy, J. L., Abrams, C. K., and Scherer, S. S. (2008). Gap junctions couple astrocytes and oligodendrocytes. J Mol Neurosci 35, 101–116.CrossRefGoogle ScholarPubMed
Ota, K., Matsui, M., Milford, E., Mackin, G., Weiner, H., and Hafler, D. (1990). T-cell recognition of an immunodominant myelin basic protein epitope in multiple sclerosis. Nature 346, 183–187.CrossRefGoogle ScholarPubMed
Ozawa, K., Suchanek, G., Breitschopf, H., et al. (1994). Patterns of oligodendroglia pathology in multiple sclerosis. Brain 117, 1311–1322.CrossRefGoogle ScholarPubMed
Padgett, B. L. and Walker, D. L. (1973). Prevalence of antibodies in human sera against JC virus, an isolate from a case of progressive multifocal leukoencephalopathy. J Infect Dis 127, 467–470.CrossRefGoogle ScholarPubMed
Padovani-Claudio, D., Lui, L., Ransohoff, R. M., and Miller, R. H. (2006). Alterations in the oligodendrocyte lineage, myelin and white matter in adult mice lacking the chemokine receptor CXCR2. Glia 54, 471–483.CrossRefGoogle ScholarPubMed
Pan, B., Fromholt, S. E., Hess, E. J., et al. (2005). Myelin-associated glycoprotein and complementary axonal ligands, gangliosides, mediate axon stability in the CNS and PNS: neuropathology and behavioral deficits in single- and double-null mice. Exp Neurol 195, 208–217.CrossRefGoogle ScholarPubMed
Pang, Y., Cai, Z., and Rhodes, P. G. (2005). Effect of tumor necrosis factor-alpha on developing optic nerve oligodendrocytes in culture. J Neurosci Res 80, 226–234.CrossRefGoogle ScholarPubMed
Panitch, H. S., Hirsch, R. L., Haley, A. S., and Johnson, K. P. (1987). Exacerbations of multiple sclerosis in patients treated with gamma interferon. Lancet 1 (8538), 893–895.CrossRefGoogle ScholarPubMed
Papastefanaki, F., Chen, J., Lavdas, A. A., Thomaidou, D., Schachner, M., and Matsas, R. (2007). Grafts of Schwann cells engineered to express PSA-NCAM promote functional recovery after spinal cord injury. Brain 130, 2159–2174.CrossRefGoogle ScholarPubMed
Park, E., Liu, Y., and Fehlings, M. G. (2003). Changes in glial cell white matter AMPA receptor expression after spinal cord injury and relationship to apoptotic cell death. Exp Neurol 182, 35–48.CrossRefGoogle ScholarPubMed
Park, H. J., Lee, P. H., Bang, O. Y., Lee, G., and Ahn, Y. H. (2008). Mesenchymal stem cells therapy exerts neuroprotection in a progressive animal model of Parkinson's disease. J Neurochem 107, 141–151.CrossRefGoogle Scholar
Park, J. B., Yiu, G., Kaneko, S., et al. (2005). A TNF receptor family member, TROY, is a coreceptor with Nogo receptor in mediating the inhibitory activity of myelin inhibitors. Neuron 45, 345–351.CrossRefGoogle ScholarPubMed
Park, S. K., Miller, R., Krane, I., and Vartanian, T. (2001). The erbB2 gene is required for the development of terminally differentiated spinal cord oligodendrocytes. J Cell Biol 154, 1245–1258.CrossRefGoogle ScholarPubMed
Parker, R. and Sheth, U. (2007). P bodies and the control of mRNA translation and degradation. Mol Cell 25, 635–646.CrossRefGoogle ScholarPubMed
Parkinson, D. B., Bhaskaran, A., Arthur-Farraj, P., et al. (2008). c-Jun is a negative regulator of myelination. J Cell Biol 181, 625–637.CrossRefGoogle ScholarPubMed
Parmantier, E., Cabon, F., Braun, C., D'Urso, D., Muller, H. W., and Zalc, B. (1995). Peripheral myelin protein-22 is expressed in rat and mouse brain and spinal cord motoneurons. Eur J Neurosci 7, 1080–1088.CrossRefGoogle ScholarPubMed
Parpura, V., Basarsky, T. A., Liu, F., Jeftinija, K., Jeftinija, S., and Haydon, P. G. (1994). Glutamate-mediated astrocyte-neuron signalling. Nature 369, 744–747.CrossRefGoogle ScholarPubMed
Patani, R., Balaratnam, M., Vora, A., and Reynolds, R. (2007). Remyelination can be extensive in multiple sclerosis despite a long disease course. Neuropathol Appl Neurobiol 33, 277–287.CrossRefGoogle ScholarPubMed
Patneau, D. K., Wright, P. W., Winters, C., Mayer, M. L., and Gallo, V. (1994). Glial cells of the oligodendrocyte lineage express both kainate- and AMPA-preferring subtypes of glutamate receptor. Neuron 12, 357–371.CrossRefGoogle ScholarPubMed
Patrikios, P., Stadelmann, C., Kutzelnigg, A., et al. (2006). Remyelination is extensive in a subset of multiple sclerosis patients. Brain 129, 3165–3172.CrossRefGoogle Scholar
Paty, D. W. and Li, D. K. (1993). Interferon beta-1b is effective in relapsing-remitting multiple sclerosis. II. MRI analysis results of a multicenter, randomized, double-blind, placebo-controlled trial. UBC MS/MRI Study Group and the IFNB Multiple Sclerosis Study Group. Neurology 43, 662–667.CrossRefGoogle Scholar
Paul, F., Jarius, S., Aktas, O., et al. (2007). Antibody to aquaporin 4 in the diagnosis of neuromyelitis optica. PLoS Med 4, e133.CrossRefGoogle Scholar
Pearse, D. D., Pereira, F. C., Marcillo, A. E., et al. (2004). cAMP and Schwann cells promote axonal growth and functional recovery after spinal cord injury. Nat Med 10, 610–616.CrossRefGoogle ScholarPubMed
Peebles, D. M., Miller, S., Newman, J. P., Scott, R., and Hanson, M. A. (2003). The effect of systemic administration of lipopolysaccharide on cerebral haemodynamics and oxygenation in the 0.65 gestation ovine fetus in utero. Br J Obstet Gynaecol 110, 735–743.CrossRefGoogle ScholarPubMed
Peirce, T. R., Bray, N. J., Williams, N. M., et al. (2006). Convergent evidence for 2′,3′-cyclic nucleotide 3′-phosphodiesterase as a possible susceptibility gene for schizophrenia. Arch Gen Psychiatry 63, 18–24.CrossRefGoogle Scholar
Pellissier, F., Gerber, A., Bauer, C., Ballivet, M., and Ossipow, V. (2007). The adhesion molecule Necl-3/SynCAM-2 localizes to myelinated axons, binds to oligodendrocytes and promotes cell adhesion. BMC Neurosci 8, 90.CrossRefGoogle ScholarPubMed
Pellkofer, H., Schubart, A. S., Hoftberger, R., et al. (2004). Modelling paraneoplastic CNS disease: T-cells specific for the onconeuronal antigen PNMA1 mediate autoimmune encephalomyelitis in the rat. Brain 127, 1822–1830.CrossRefGoogle ScholarPubMed
Pende, M., Holtzclaw, L. A., Curtis, J. L., Russell, J. T., and Gallo, V. (1994). Glutamate regulates intracellular calcium and gene expression in oligodendrocyte progenitors through the activation of DL-α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptors. Proc Natl Acad Sci USA 91, 3215–3219.CrossRefGoogle ScholarPubMed
Penderis, J., Shields, S. A., and Franklin, R. J. (2003). Impaired remyelination and depletion of oligodendrocyte progenitors does not occur following repeated episodes of focal demyelination in the rat central nervous system. Brain 126, 1382–1391.CrossRefGoogle Scholar
Penfield, W. (1924). Oligodendroglia and its relation to classical neuroglia. Brain 47, 430–452.CrossRefGoogle Scholar
Peress, N. S., Perillo, E., and Seidman, R. J. (1996). Glial transforming growth factor (TGF)-beta isotypes in multiple sclerosis: differential glial expression of TGF-beta 1, 2 and 3 isotypes in multiple sclerosis. J Neuroimmunol 71, 115–123.CrossRefGoogle ScholarPubMed
Perez Villages, E. M., Olivier, C., Spassky, N., et al. (1999). Early specification of oligodendrocytes in the chick embryonic brain. Dev Biol 216, 98–113.CrossRefGoogle Scholar
Perier, O. and Gregoire, A. (1965). Electron microscopic features of multiple sclerosis lesions. Brain 88, 937–952.CrossRefGoogle ScholarPubMed
Perkins, D. O., Jeffries, C. D., Jarskog, L. F., et al. (2007). microRNA expression in the prefrontal cortex of individuals with schizophrenia and schizoaffective disorder. Genome Biol 8, R27.CrossRefGoogle ScholarPubMed
Perron, H., Jouvin-Marche, E., Michel, M., et al. (2001). Multiple sclerosis retrovirus particles and recombinant envelope trigger an abnormal immune response in vitro, by inducing polyclonal Vbeta16 T-lymphocyte activation. Virology 287, 321–332.CrossRefGoogle ScholarPubMed
Perron, H., Lazarini, F., Ruprecht, K., et al. (2005). Human endogenous retrovirus (HERV)-W ENV and GAG proteins: physiological expression in human brain and pathophysiological modulation in multiple sclerosis lesions. J Neurovirol 11, 23–33.CrossRefGoogle ScholarPubMed
Perry, V. H., Brown, M. C., and Andersson, P. B. (1993). Macrophage responses to central and peripheral nerve injury. Adv Neurol 59, 309–314.Google ScholarPubMed
Peters, A. and Sethares, C. (2002). Aging and the myelinated fibers in prefrontal cortex and corpus callosum of the monkey. J Comp Neurol 442, 277–291.CrossRefGoogle ScholarPubMed
Peters, A. and Sethares, C. (2003). Is there remyelination during aging of the primate central nervous system? J Comp Neurol 460, 238–254.CrossRefGoogle ScholarPubMed
Peters, A. and Sethares, C. (2004). Oligodendrocytes, their progenitors and other neuroglial cells in the aging primate cerebral cortex. Cereb Cortex 14, 995–1007.CrossRefGoogle ScholarPubMed
Peters, A., Moss, M. B., and Sethares, C. (2000). Effects of aging on myelinated nerve fibers in monkey primary visual cortex. J Comp Neurol 419, 364–376.3.0.CO;2-R>CrossRefGoogle ScholarPubMed
Peters, A., Palay, S. L., and Webster, H. D. (1991). The Fine Structure of the Nervous System: Neurons and their Supporting Cells. (New York: Oxford University Press.)Google Scholar
Peters, A., Rosene, D. L., Moss, M. B., et al. (1996). Neurobiological bases of age-related cognitive decline in the rhesus monkey. J Neuropathol Exp Neurol 55, 861–874.CrossRefGoogle ScholarPubMed
Peters, A., Sethares, C., and Killiany, R. J. (2001). Effects of age on the thickness of myelin sheaths in monkey primary visual cortex. J Comp Neurol 435, 241–248.CrossRefGoogle ScholarPubMed
Pfeiffer, S. E., Warrington, A. E., and Bansal, R. (1993). The oligodendrocyte and its many cellular processes. Trends Cell Biol 3, 191–197.CrossRefGoogle ScholarPubMed
Picard-Riera, N., Decker, L., Delarasse, C., et al. (2002). Experimental autoimmune encephalomyelitis mobilizes neural progenitors from the subventricular zone to undergo oligodendrogenesis in adult mice. Proc Natl Acad Sci USA 99, 13211–13216.CrossRefGoogle ScholarPubMed
Piddlesden, S. J. and Morgan, B. P. (1993). Killing of rat glial cells by complement: deficiency of the rat analogue of CD59 is the cause of oligodendrocyte susceptibility to lysis. J Neuroimmunol 48, 169–176.CrossRefGoogle ScholarPubMed
Pierson, C. R., Folkerth, R. D., Billiards, S. S., et al. (2007). Gray matter injury associated with periventricular leukomalacia in the premature infant. Acta Neuropathol 114, 619–631.CrossRefGoogle ScholarPubMed
Pines, G., Danbolt, N. C., Bjørås, M., et al. (1992). Cloning and expression of a rat brain L-glutamate transporter. Nature 360, 464–467.CrossRefGoogle ScholarPubMed
Pitt, D., Nagelmeier, I. E., Wilson, H. C., and Raine, C. S. (2003). Glutamate uptake by oligodendrocytes: implications for excitotoxicity in multiple sclerosis. Neurology 61, 1113–1120.CrossRefGoogle ScholarPubMed
Pitt, D., Werner, P., and Raine, C. S. (2000). Glutamate excitotoxicity in a model of multiple sclerosis. Nature Med 6, 67–70.CrossRefGoogle Scholar
Pizzi, M., Sarnico, I., Boroni, F., et al. (2004). Prevention of neuron and oligodendrocyte degeneration by interleukin-6 (IL-6) and IL-6 receptor/IL-6 fusion protein in organotypic hippocampal slices. Mol Cell Neurosci 25, 301–311.CrossRefGoogle ScholarPubMed
Pluchino, S., Quattrini, A., Brambilla, E., et al. (2003). Injection of adult neurospheres induces recovery in a chronic model of multiple sclerosis. Nature 422, 688–694.CrossRefGoogle Scholar
Pluchino, S., Zanotti, L., Deleidi, M., and Martino, G. (2005a). Neural stem cells and their use as therapeutic tool in neurological disorders. Brain Res Brain Res Rev 48, 211–219.CrossRefGoogle ScholarPubMed
Pluchino, S., Zanotti, L., Rossi, B., et al. (2005b). Neurosphere-derived multipotent precursors promote neuroprotection by an immunomodulatory mechanism. Nature 436, 266–271.CrossRefGoogle ScholarPubMed
Poliak, S. and Peles, E. (2003). The local differentiation of myelinated axons at nodes of Ranvier. Nat Rev Neurosci 4, 968–980.CrossRefGoogle ScholarPubMed
Poliak, S., Gollan, L., Martinez, R., et al. (1999). Caspr2, a new member of the neurexin superfamily, is localized at the juxtaparanodes of myelinated axons and associates with K+ channels. Neuron 24, 1037–1047.CrossRefGoogle ScholarPubMed
Polito, A. and Reynolds, R. (2005). NG2-expressing cells as oligodendrocyte progenitors in the normal and demyelinated adult central nervous system. J Anat 207, 707–716.CrossRefGoogle ScholarPubMed
Poltorak, M., Sadoul, R., Keilhauer, G., Landa, C., Fahrig, T., and Schachner, M. (1987). Myelin-associated glycoprotein, a member of the L2/HNK-1 family of neural cell adhesion molecules, is involved in neuron-oligodendrocyte and oligodendrocyte-oligodendrocyte interaction. J Cell Biol 105, 1893–1899.CrossRefGoogle ScholarPubMed
Popko, B. and Baerwald, K. D. (1999). Oligodendroglial response to the immune cytokine interferon gamma. Neurochem Res 24, 331–338.CrossRefGoogle ScholarPubMed
Pot, C., Simonen, M., Weinmann, O., et al. (2002). Nogo-A expressed in Schwann cells impairs axonal regeneration after peripheral nerve injury. J Cell Biol 159, 29–35.CrossRefGoogle ScholarPubMed
Pouly, S., Becher, B., Blain, M., and Antel, J. P. (2000). Interferon-γ modulates human oligodendrocyte susceptibility to Fas-mediated apoptosis. J Neuropathol Exp Neurol 59, 280–286.CrossRefGoogle ScholarPubMed
Prestoz, L., Chatzopoulou, E., Lemkine, G., et al. (2004). Control of axonophilic migration of oligodendrocyte precursor cells by Eph-ephrin interaction. Neuron Glia Biol 1, 73–83.CrossRefGoogle ScholarPubMed
Prineas, J. W. and Connell, F. (1979). Remyelination in multiple sclerosis. Ann Neurol 5, 22–31.CrossRefGoogle ScholarPubMed
Prineas, J. W. and Graham, J. S. (1981). Multiple sclerosis: capping of surface immunoglobulin G on macrophages engaged in myelin breakdown. Ann Neurol 10, 149–158.CrossRefGoogle Scholar
Prineas, J. W., Barnard, R. O., Kwon, E. E., Sharer, L. R., and Cho, E. S. (1993a). Multiple sclerosis: remyelination of nascent lesions. Ann Neurol 33, 137–151.CrossRefGoogle ScholarPubMed
Prineas, J. W., Barnard, R. O., Revesz, T., Kwon, E. E., Sharer, L., and Cho, E. S. (1993b). Multiple sclerosis. Pathology of recurrent lesions. Brain 116, 681–693.CrossRefGoogle ScholarPubMed
Pringle, N. P. and Richardson, W. D. (1993). A singularity of PDGF alpha-receptor expression in the dorsoventral axis of the neural tube may define the origin of the oligodendrocyte lineage. Development 117, 525–533.Google ScholarPubMed
Pringle, N. P., Mudhar, H. S., Collarini, E. J., and Richardson, W. D. (1992). PDGF receptors in the rat CNS: during late neurogenesis, PDGF alpha-receptor expression appears to be restricted to glial cells of the oligodendrocyte lineage. Development 115, 535–551.Google ScholarPubMed
Pringle, N. P., Yu, W. P., Guthrie, S., et al. (1996). Determination of neuroepithelial cell fate: induction of the oligodendrocyte lineage by ventral midline cells and sonic hedgehog. Dev Biol 177, 30–42.CrossRefGoogle ScholarPubMed
Prinjha, R., Moore, S. E., Vinson, M., et al. (2000). Inhibitor of neurite outgrowth in humans. Nature 403, 383–384.CrossRefGoogle ScholarPubMed
Privat, A., Jacque, C., Bourre, J. M., Dupouey, P., and Baumann, N. (1979). Absence of the major dense line in myelin of the mutant mouse “shiverer”. Neurosci Lett 12, 107–112.CrossRefGoogle ScholarPubMed
Proudfoot, A. E. (2002). Chemokine receptors: multifaceted therapeutic targets. Nat Rev Immunol 2, 106–115.CrossRefGoogle ScholarPubMed
Puchalski, R. B., Louis, J. C., Brose, N., et al. (1994). Selective RNA editing and subunit assembly of native glutamate receptors. Neuron 13, 131–147.CrossRefGoogle ScholarPubMed
Puckett, C., Hudson, L., Ono, K., et al. (1987). Myelin-specific proteolipid protein is expressed in myelinating Schwann cells but is not incorporated into myelin sheaths. J Neurosci Res 18, 511–518.CrossRefGoogle Scholar
Pujol, J., Soriano-Mas, C., Ortiz, H., Sebastian-Galles, N., Losilla, J. M., and Deus, J. (2006). Myelination of language-related areas in the developing brain. Neurology 66, 339–343.CrossRefGoogle ScholarPubMed
Rabchevsky, A. G., Sullivan, P. G., and Scheff, S. W. (2007). Temporal-spatial dynamics in oligodendrocyte and glial progenitor cell numbers throughout ventrolateral white matter following contusion spinal cord injury. Glia 55, 831–843.CrossRefGoogle ScholarPubMed
Racke, M. K. (2008). The role of B cells in multiple sclerosis: rationale for B-cell-targeted therapies. Curr Opin Neurol 21 Suppl 1, S9–S18.CrossRefGoogle ScholarPubMed
Radi, R., Beckman, J. S., Bush, K. M., and Freeman, B. A. (1991). Peroxynitrite-induced membrane lipid peroxidation: the cytotoxic potential of superoxide and nitric oxide. Arch Biochem Biophys 288, 481–487.CrossRefGoogle ScholarPubMed
Raff, M. C. (1989). Glial cell diversification in the rat optic nerve. Science 243, 1450–1455.CrossRefGoogle ScholarPubMed
Raff, M. C., Miller, R. H., and Noble, M. (1983). A glial progenitor cell that develops in vitro into an astrocyte or an oligodendrocyte depending on culture medium. Nature 303, 390–396.CrossRefGoogle ScholarPubMed
Raff, M. C., Mirsky, R., Fields, K. L., et al. (1978). Galactocerebroside is a specific cell-surface antigenic marker for oligodendrocytes in culture. Nature 274, 813–816.CrossRefGoogle ScholarPubMed
Raine, C. S. (1991). Multiple sclerosis: a pivotal role for the T cell in lesion development. Neuropathol Appl Neurobiol 17, 265–274.CrossRefGoogle Scholar
Raine, C. S. and Cross, A. H. (1989). Axonal dystrophy as a consequence of long-term demyelination. Lab Invest 60, 714–725.Google ScholarPubMed
Raine, C. S. and Wu, E. (1993). Multiple sclerosis: remyelination in acute lesions. J Neuropathol Exp Neurol 52, 199–204.CrossRefGoogle ScholarPubMed
Raine, C. S., Scheinberg, L., and Waltz, J. M. (1981). Multiple sclerosis. Oligodendrocyte survival and proliferation in an active established lesion. Lab Invest 45, 534–546.Google Scholar
Raisman, G. (2004). Olfactory ensheathing cells and repair of brain and spinal cord injuries. Cloning Stem Cells 6, 364–368.CrossRefGoogle ScholarPubMed
Ranscht, B. and Dours, M. T. (1988). Sequence of contactin, a 130-kD glycoprotein concentrated in areas of interneuronal contact, defines a new member of the immunoglobulin supergene family in the nervous system. J Cell Biol 107, 1561–1573.CrossRefGoogle Scholar
Ranvier, L. A. (1878). Leçons sur l'Histologie du Système Nerveux. (Paris: Librarie F. Savy.)Google Scholar
Rasband, M. N., Park, E. W., Zhen, D., et al. (2002). Clustering of neuronal potassium channels is independent of their interaction with PSD-95. J Cell Biol 159, 663–672.CrossRefGoogle ScholarPubMed
Rasband, M. N., Tayler, J., Kaga, Y., et al. (2005). CNP is required for maintenance of axon-glia interactions at nodes of Ranvier in the CNS. Glia 50, 86–90.CrossRefGoogle ScholarPubMed
Rasband, M. N., Trimmer, J. S., Schwarz, T. L., et al. (1998). Potassium channel distribution, clustering, and function in remyelinating rat axons. J Neurosci 18, 36–47.CrossRefGoogle ScholarPubMed
Rash, J. E., Yasumura, T., Davidson, K. G., Furman, C. S., Dudek, F. E., and Nagy, J. I. (2001). Identification of cells expressing Cx43, Cx30, Cx26, Cx32 and Cx36 in gap junctions of rat brain and spinal cord. Cell Commun Adhes 8, 315–320.CrossRefGoogle ScholarPubMed
Rauer, S., Euler, B., Reindl, M., and Berger, T. (2006). Antimyelin antibodies and the risk of relapse in patients with a primary demyelinating event. J Neurol Neurosurg Psychiatry 77, 739–742.CrossRefGoogle ScholarPubMed
Readhead, C., Popko, B., Takahashi, N., et al. (1987). Expression of a myelin basic protein gene in shiverer transgenic mice: correction of the dysmyelinating phenotype. Cell 48, 703–712.CrossRefGoogle ScholarPubMed
Readhead, C., Schneider, A., Griffiths, I., and Nave, K.-A. (1994). Premature arrest of myelin formation in transgenic mice with increased proteolipid protein gene dosage. Neuron 12, 583–595.CrossRefGoogle ScholarPubMed
Redwine, J. M., Buchmeier, M. J., and Evans, C. F. (2001). In vivo expression of major histocompatibility complex molecules on oligodendrocytes and neurons during viral infection. Am J Pathol 159, 1219–1224.CrossRefGoogle ScholarPubMed
Reier, J. P. (1988). The glial scar: its bearing on axonal elongation and transplantation approaches to CNS repair. In Functional Recovery in Neurological Disease, Waxman, S. G., ed. (New York: Raven), pp. 87–138.Google Scholar
Relvas, J. B., Setzu, A., Baron, W., et al. (2001). Expression of dominant-negative and chimeric subunits reveals an essential role for beta1 integrin during myelination. Curr Biol 11, 1039–1043.CrossRefGoogle ScholarPubMed
Remahl, S. and Hildebrand, C. (1990). Relation between axons and oligodendroglial cells during initial myelination. I. The glial unit. J Neurocytol 19, 313–328.CrossRefGoogle ScholarPubMed
Reynolds, B. A. and Weiss, S. (1992). Generation of neurons and astrocytes from isolated cells of the adult mammalian central nervous system. Science 255, 1707–1710.CrossRefGoogle ScholarPubMed
Reynolds, B. A., Tetzlaff, W., and Weiss, S. (1992). A multipotent EGF-responsive striatal embryonic progenitor cell produces neurons and astrocytes. J Neurosci 12, 4565–4574.CrossRefGoogle ScholarPubMed
Reynolds, R. and Hardy, R. (1997). Oligodendroglial progenitors labeled with the O4 antibody persist in the adult rat cerebral cortex in vivo. J Neurosci Res 47, 455–470.3.0.CO;2-G>CrossRefGoogle ScholarPubMed
Reynolds, R., Dawson, M., Papadopoulos, D., et al. (2002). The response of NG2-expressing oligodendrocyte progenitors to demyelination in MOG-EAE and MS. J Neurocytol 31, 523–536.CrossRefGoogle ScholarPubMed
Ricci-Vitiani, L., Conticello, C., Zeuner, A., and Maria, R. (2000). CD95/CD95L interactions and their role in autoimmunity. Apoptosis 5, 419–424.CrossRefGoogle ScholarPubMed
Rice, C. M. and Scolding, N. J. (2004). Adult stem cells – reprogramming neurological repair? Lancet 364, 193–199.CrossRefGoogle ScholarPubMed
Rice, C. M., and Scolding, N. J. (2008). Autologous bone marrow stem cells – properties and advantages. J Neurol Sci 15, 59–62.CrossRefGoogle Scholar
Rice, C. M., Whone, A. L., Marks, D. I., Butler, S. B., Brooks, D. J., and Scolding, N. J. (2007). A safety and feasibility study of intravenous autologous bone marrow stem cells in multiple sclerosis. J Neurol Neurosurg Psychiatry 78, 1014–1038.Google Scholar
Rice, G. P., Hartung, H. P., and Calabresi, P. A. (2005). Anti-alpha4 integrin therapy for multiple sclerosis: mechanisms and rationale. Neurology 64, 1336–1342.CrossRefGoogle ScholarPubMed
Richardson, P. M., McGuinness, U. M., and Aguayo, A. J. (1980). Axons from CNS neurons regenerate into PNS grafts. Nature 284, 264–265.CrossRefGoogle ScholarPubMed
Richardson, W. D., Kessaris, N., and Pringle, N. (2006). Oligodendrocyte wars. Nat Rev Neurosci 7, 11–18.CrossRefGoogle ScholarPubMed
Richardson, W. D., Pringle, N., Mosley, M. J., Westermark, B., and Dubois-Dalcq, M. (1988). A role for platelet-derived growth factor in normal gliogenesis in the central nervous system. Cell 53, 309–319.CrossRefGoogle ScholarPubMed
Richter-Landsberg, C. and Gorath, M. (1999). Developmental regulation of alternatively spliced isoforms of mRNA encoding MAP2 and tau in rat brain oligodendrocytes during culture maturation. J Neurosci Res 56, 259–270.3.0.CO;2-N>CrossRefGoogle ScholarPubMed
Riethmacher, D., Sonnenberg-Riethmacher, E., Brinkmann, V., Yamaai, T., Lewin, G. R., and Birchmeier, C. (1997). Severe neuropathies in mice with targeted mutations in the ErbB3 receptor. Nature 389, 725–730.CrossRefGoogle Scholar
Rincon-Orozco, B., Kunzmann, V., Wrobel, P., Kabelitz, D., Steinle, A., and Herrmann, T. (2005). Activation of V gamma 9V delta 2 T cells by NKG2D. J Immunol 175, 2144–2151.CrossRefGoogle ScholarPubMed
Rios, J. C., Melendez-Vasquez, C. V., Einheber, S., et al. (2000). Contactin-associated protein (Caspr) and contactin form a complex that is targeted to the paranodal junctions during myelination. J Neurosci 20, 8354–8364.CrossRefGoogle Scholar
Rivera, F. J., Couillard-Despres, S., Pedre, X., et al. (2006). Mesenchymal stem cells instruct oligodendrogenic fate decision on adult neural stem cells. Stem Cells 24, 2209–2219.CrossRefGoogle ScholarPubMed
Rivers, L. E., Young, K. M., Rizzi, M., et al. (2008). PDGFRA/NG2 glia generate myelinating oligodendrocytes and piriform projection neurons in adult mice. Nat Neurosci 11, 1392–1401.CrossRefGoogle ScholarPubMed
Robinson, S., Tani, M., Strieter, R. M., Ransohoff, R. M., and Miller, R. H. (1998). The chemokine growth-regulated oncogene-alpha promotes spinal cord oligodendrocyte precursor proliferation. J Neurosci 18, 10457–10463.CrossRefGoogle ScholarPubMed
Rodriguez, M. (2003). A function of myelin is to protect axons from subsequent injury: implications for deficits in multiple sclerosis. Brain 126, 751–752.CrossRefGoogle ScholarPubMed
Rodriguez, M., Scheithauer, B. W., Forbes, G., and Kelly, P. J. (1993). Oligodendrocyte injury is an early event in lesions of multiple sclerosis. Mayo Clin Proc 68, 627–636.CrossRefGoogle ScholarPubMed
Roettger, V. and Lipton, P. (1996). Mechanism of glutamate release from rat hippocampal slices during in vitro ischemia. Neuroscience 75, 677–685.CrossRefGoogle ScholarPubMed
Rosenberg, P. A. and Aizenman, E. (1989). Hundred-fold increase in neuronal vulnerability to glutamate toxicity in astrocyte-poor cultures of rat cerebral cortex. Neurosci Lett 103, 162–168.CrossRefGoogle ScholarPubMed
Rosenberg, P. A., Amin, S., and Leitner, M. (1992). Glutamate uptake disguises neurotoxic potency of glutamate agonists in cerebral cortex in dissociated cell culture. J Neurosci 12, 56–61.CrossRefGoogle ScholarPubMed
Rosenberg, P. A., Dai, W. M., Gan, X. D., et al. (2003). Mature myelin basic protein-expressing oligodendrocytes are insensitive to kainate toxicity. J Neurosci Res 71, 237–245.CrossRefGoogle ScholarPubMed
Rosenbluth, J. (1980a). Central myelin in the mouse mutant shiverer. J Comp Neurol 194, 639–648.CrossRefGoogle ScholarPubMed
Rosenbluth, J. (1980b). Peripheral myelin in the mouse mutant Shiverer. J Comp Neurol 193, 729–739.CrossRefGoogle ScholarPubMed
Rosenbluth, J., Schiff, R., and Lam, P. (2009). Effects of osmolality on PLP-null myelin structure: implications re axon damage. Brain Res 1253, 191–197.CrossRefGoogle ScholarPubMed
Rossi, D. and Zlotnik, A. (2000). The biology of chemokines and their receptors. Annu Rev Immunol 18, 217–242.CrossRefGoogle ScholarPubMed
Rossi, D. J., Oshima, T., and Attwell, D. (2000). Glutamate release in severe brain ischaemia is mainly by reversed uptake. Nature 403, 316–321.CrossRefGoogle ScholarPubMed
Roth, A. D., Ivanova, A., and Colman, D. R. (2006). New observations on the compact myelin proteome. Neuron Glia Biol 2, 15–21.CrossRefGoogle ScholarPubMed
Rothman, R. H., and Simeone, F. A. (1992). The Spine. (Philadelphia, PA: Saunders.)Google Scholar
Rousset, C. I., Kassem, J., Olivier, P., Chalon, S., Gressens, P., and Saliba, E. (2008). Antenatal bacterial endotoxin sensitizes the immature rat brain to postnatal excitotoxic injury. J Neuropathol Exp Neurol 67, 994–1000.CrossRefGoogle ScholarPubMed
Rowitch, D. H. (2004). Glial specification in the vertebrate neural tube. Nat Rev Neurosci 5, 409–419.CrossRefGoogle ScholarPubMed
Rowitch, D. H., Lu, Q. R., Kessaris, N., and Richardson, W. D. (2002). An “oligarchy” rules neural development. Trends Neurosci 25, 417–422.CrossRefGoogle Scholar
Roy, K., Murtie, J. C., El Khodor, B. F., et al. (2007). Loss of erbB signaling in oligodendrocytes alters myelin and dopaminergic function, a potential mechanism for neuropsychiatric disorders. Proc Natl Acad Sci USA 104, 8131–8136.CrossRefGoogle ScholarPubMed
Roy, N. S., Cleren, C., Singh, S. K., Yang, L., Beal, M. F., and Goldman, S. A. (2006). Functional engraftment of human ES cell-derived dopaminergic neurons enriched by coculture with telomerase-immortalized midbrain astrocytes. Nat Med 12, 1259–1268.CrossRefGoogle ScholarPubMed
Ruijs, T. C., Freedman, M. S., Grenier, Y. G., Olivier, A., and Antel, J. P. (1990). Human oligodendrocytes are susceptible to cytolysis by major histocompatibility complex class I-restricted lymphocytes. J Neuroimmunol 27, 89–97.CrossRefGoogle ScholarPubMed
Rutkowski, J. L., Kirk, C. J., Lerner, M. A., and Tennekoon, G. I. (1995). Purification and expansion of human schwann cells in vitro. Nature Med 1, 80–83.CrossRefGoogle ScholarPubMed
Saher, G., Brugger, B., Lappe-Siefke, C., et al. (2005). High cholesterol level is essential for myelin membrane growth. Nat Neurosci 8, 468–475.CrossRefGoogle ScholarPubMed
Saikali, P., Antel, J. P., Newcombe, J., et al. (2007). NKG2D-mediated cytotoxicity toward oligodendrocytes suggests a mechanism for tissue injury in multiple sclerosis. J Neurosci 27, 1220–1228.CrossRefGoogle ScholarPubMed
Salter, M. G. and Fern, R. (2005). NMDA receptors are expressed in developing oligodendrocyte processes and mediate injury. Nature 438, 1167–1171.CrossRefGoogle ScholarPubMed
Salviati, L., Trevisson, E., Baldoin, M. C., et al. (2007). A novel deletion in the GJA12 gene causes Pelizaeus-Merzbacher-like disease. Neurogenetics 8, 57–60.CrossRefGoogle ScholarPubMed
Salzer, J. L. (2003). Polarized domains of myelinated axons. Neuron 40, 297–318.CrossRefGoogle ScholarPubMed
Salzer, J. L., Brophy, P. J., and Peles, E. (2008). Molecular domains of myelinated axons in the peripheral nervous system. Glia 56, 1532–1540.CrossRefGoogle ScholarPubMed
Salzer, J. L., Holmes, W. P., and Colman, D. R. (1987). The amino acid sequences of the myelin-associated glycoproteins: homology to the immunoglobulin gene superfamily. J Cell Biol 104, 957–965.CrossRefGoogle ScholarPubMed
Sandler, A. N. and Tator, C. H. (1976). Effect of acute spinal cord compression injury on regional spinal cord blood flow in primates. J Neurosurg 45, 660–676.CrossRefGoogle ScholarPubMed
Sasaki, M., Honmou, O., Akiyama, Y., Uede, T., Hashi, K., and Kocsis, J. D. (2001). Transplantation of an acutely isolated bone marrow fraction repairs demyelinated adult rat spinal cord axons. Glia 35, 26–34.CrossRefGoogle ScholarPubMed
Sasaki, Y. F., Rothe, T., Premkumar, L. S., et al. (2002). Characterization and comparison of the NR3A subunit of the NMDA receptor in recombinant systems and primary cortical neurons. J Neurophysiol 87, 2052–2063.CrossRefGoogle ScholarPubMed
Satoh, J., Kastrukoff, L. F., and Kim, S. U. (1991). Cytokine-induced expression of intercellular adhesion molecule-1 (ICAM-1) in cultured human oligodendrocytes and astrocytes. J Neuropathol Exp Neurol 50, 215–226.CrossRefGoogle ScholarPubMed
Savio, T. and Schwab, M. E. (1990). Lesioned corticospinal tract axons regenerate in myelin-free rat spinal cord. Proc Natl Acad Sci USA 87, 4130–4133.CrossRefGoogle ScholarPubMed
Sawada, M., Itoh, Y., Suzumura, A., and Marunouchi, T. (1993). Expression of cytokine receptors in cultured neuronal and glial cells. Neurosci Lett 160, 131–134.CrossRefGoogle ScholarPubMed
Schabitz, W. R., Li, F., and Fisher, M. (2000). The N-methyl-d-aspartate antagonist CNS 1102 protects cerebral gray and white matter from ischemic injury following temporary focal ischemia in rats. Stroke 31, 1709–1714.CrossRefGoogle ScholarPubMed
Schachner, M. and Bartsch, U. (2000). Multiple functions of the myelin-associated glycoprotein MAG (siglec-4a) in formation and maintenance of myelin. Glia 29, 154–165.3.0.CO;2-3>CrossRefGoogle ScholarPubMed
Schaefer, A., O'Carroll, D., Tan, C. L., et al. (2007). Cerebellar neurodegeneration in the absence of microRNAs. J Exp Med 204, 1553–1558.CrossRefGoogle ScholarPubMed
Schaeren-Wiemers, N., Schaefer, C., Valenzuela, D. M., Yancopoulos, G. D., and Schwab, M. E. (1995). Identification of new oligodendrocyte- and myelin-specific genes by a differential screening approach. J Neurochem 65, 10–22.CrossRefGoogle ScholarPubMed
Schafer, D. P. and Rasband, M. N. (2006). Glial regulation of the axonal membrane at nodes of Ranvier. Curr Opin Neurobiol 16, 508–514.CrossRefGoogle ScholarPubMed
Schafer, M., Fruttiger, M., Montag, D., Schachner, M., and Martini, R. (1996). Disruption of the gene for the myelin-associated glycoprotein improves axonal regrowth along myelin in C57BL/Wlds mice. Neuron 16, 1107–1113.CrossRefGoogle ScholarPubMed
Scheinman, R. I., Cogswell, P. C., Lofquist, A. K., and Baldwin, A. S.. (1995). Role of transcriptional activation of I kappa B alpha in mediation of immunosuppression by glucocorticoids. Science 270, 283–286.CrossRefGoogle ScholarPubMed
Scheld, W. M., Whitley, R. J., and Marra, C. M. eds. (2004). Infections of the Central Nervous System, 3rd edn. (Philadelphia, PA: Lippincott Williams & Wilkins.)Google Scholar
Scherer, S. S. and Arroyo, E. J. (2002). Recent progress on the molecular organization of myelinated axons. J Peripher Nerv Syst 7, 1–12.CrossRefGoogle ScholarPubMed
Scherer, S. S. and Easter, S. S.. (1984). Degenerative and regenerative changes in the trochlear nerve of goldfish. J Neurocytol 13, 519–565.CrossRefGoogle ScholarPubMed
Scherer, S. S. and Wrabetz, L. (2008). Molecular mechanisms of inherited demyelinating neuropathies. Glia 56, 1578–1589.CrossRefGoogle ScholarPubMed
Scherer, S. S., Deschenes, S. M., Xu, Y. T., Grinspan, J. B., Fischbeck, K. H., and Paul, D. L. (1995). Connexin32 is a myelin-related protein in the PNS and CNS. J Neurosci 15, 8281–8294.CrossRefGoogle ScholarPubMed
Scherer, S. S., Vogelbacker, H. H., and Kamholz, J. (1992). Axons modulate the expression of proteolipid protein in the CNS. J Neurosci Res 32, 138–148.CrossRefGoogle ScholarPubMed
Scherer, S. S., Xu, Y. T., Nelles, E., Fischbeck, K., Willecke, K., and Bone, L. J. (1998). Connexin32-null mice develop demyelinating peripheral neuropathy. Glia 24, 8–20.3.0.CO;2-3>CrossRefGoogle ScholarPubMed
Schlag, B. D., Vondrasek, J. R., Munir, M., et al. (1998). Regulation of the glial Na+-dependent glutamate transporters by cyclic AMP analogs and neurons. Mol Pharmacol 53, 355–369.CrossRefGoogle ScholarPubMed
Schliwa, M. and Woehlke, G. (2003). Molecular motors. Nature 422, 759–765.CrossRefGoogle ScholarPubMed
Schlomann, U., Rathke-Hartlieb, S., Yamamoto, S., Jockusch, H., and Bartsch, J. W. (2000). Tumor necrosis factor alpha induces a metalloprotease-disintegrin, ADAM8 (CD 156): implications for neuron-glia interactions during neurodegeneration. J Neurosci 20, 7964–7971.CrossRefGoogle ScholarPubMed
Schmitz, T., Heep, A., Groenendaal, F., et al. (2007). Interleukin-1beta, interleukin-18, and interferon-gamma expression in the cerebrospinal fluid of premature infants with posthemorrhagic hydrocephalus – markers of white matter damage? Pediatr Res 61, 722–726.CrossRefGoogle ScholarPubMed
Schnapp, B. and Mugnaini, E. (1976). Freeze-fracture properties of central myelin in the bullfrog. Neuroscience 1, 459–467.CrossRefGoogle ScholarPubMed
Schneider, A. and Mandelkow, E. (2008). Tau-based treatment strategies in neurodegenerative diseases. Neurotherapeutics 5, 443–457.CrossRefGoogle ScholarPubMed
Schnell, L. and Schwab, M. E. (1990). Axonal regeneration in the rat spinal cord produced by an antibody against myelin-associated neurite growth inhibitors. Nature 343, 269–272.CrossRefGoogle ScholarPubMed
Scholz, J., Klein, M. C., Behrens, T. E., and Johansen-Berg, H. (2009). Training induces changes in white-matter architecture. Nat Neurosci 12, 1370–1371.CrossRefGoogle ScholarPubMed
Schonberg, D. L., Popovich, P. G., and McTigue, D. M. (2007). Oligodendrocyte generation is differentially influenced by toll-like receptor (TLR) 2 and TLR4-mediated intraspinal macrophage activation. J Neuropathol Exp Neurol 66, 1124–1135.CrossRefGoogle ScholarPubMed
Schroder, M. and Kaufman, R. J. (2005). The mammalian unfolded protein response. Annu Rev Biochem 74, 739–789.CrossRefGoogle ScholarPubMed
Schultz, R. L. and Pease, D. C. (1959). Cicatrix formation in rat cerebral cortex as revealed by electron microscopy. Am J Pathol 35, 1017–1041.Google ScholarPubMed
Schwab, C. and McGeer, P. L. (2002). Complement activated C4d immunoreactive oligodendrocytes delineate small cortical plaques in multiple sclerosis. Exp Neurol 174, 81–88.CrossRefGoogle ScholarPubMed
Schwab, M. E. (2004). Nogo and axon regeneration. Curr Opin Neurobiol 14, 118–124.CrossRefGoogle ScholarPubMed
Schwab, M. E. and Bartholdi, D. (1996). Degeneration and regeneration of axons in the lesioned spinal cord. Physiol Rev 76, 319–370.CrossRefGoogle ScholarPubMed
Schwab, M. E. and Caroni, P. (1988). Oligodendrocytes and CNS myelin are nonpermissive substrates for neurite growth and fibroblast spreading in vitro. J Neurosci 8, 2381–2393.CrossRefGoogle ScholarPubMed
Schwab, M. E. and Thoenen, H. (1985). Dissociated neurons regenerate into sciatic but not optic nerve explants in culture irrespective of neurotrophic factors. J Neurosci 5, 2415–2423.CrossRefGoogle Scholar
Schwarz, S., Knauth, M., Schwab, S., Walter-Sack, I., Bonmann, E., and Storch-Hagenlocher, B. (2000). Acute disseminated encephalomyelitis after parenteral therapy with herbal extracts: a report of two cases. J Neurol Neurosurg Psychiatry 69, 516–518.CrossRefGoogle ScholarPubMed
Schweigreiter, R., Roots, B. I., Bandtlow, C. E., and Gould, R. M. (2006). Understanding myelination through studying its evolution. Int Rev Neurobiol 73, 219–273.CrossRefGoogle ScholarPubMed
Schweitzer, J., Becker, T., Schachner, M., Nave, K. A., and Werner, H. (2006). Evolution of myelin proteolipid proteins: gene duplication in teleosts and expression pattern divergence. Mol Cell Neurosci 31, 161–177.CrossRefGoogle ScholarPubMed
Scolding, N. (2005). Stem-cell therapy: hope and hype. Lancet 365, 2073–2075.CrossRefGoogle ScholarPubMed
Scolding, N. J. and Compston, D. A. (1991). Oligodendrocyte-macrophage interactions in vitro triggered by specific antibodies. Immunology 72, 127–132.Google ScholarPubMed
Scolding, N. J. and Dubois-Dalcq, M. (2008). Moving toward remyelinating and neuroprotective therapies in MS. In Multiple Sclerosis: A Comprehensive Text, Raine, C. S., McFarland, H., and Hohlfeld, R., eds. (Philadelphia, PA: Saunders Elsevier), pp. 366–382.Google Scholar
Scolding, N. J. and Rice, C. (2008). Autologous mesenchymal bone marrow stem cells: practical considerations. J Neurol Sci 15, 111–115.CrossRefGoogle Scholar
Scolding, N. J., Franklin, R. J. M., Stevens, S., Heldin, C. H., Compston, D. A. S., and Newcombe, J. (1998a). Oligodendrocyte progenitors are present in the normal adult human CNS and in the lesions of multiple sclerosis. Brain 121, 2221–2228.CrossRefGoogle ScholarPubMed
Scolding, N. J., Jones, J., Compston, D. A. S., and Morgan, B. P. (1990). Oligodendrocyte susceptibility to injury by T-cell perforin. Immunology 70, 6–10.Google ScholarPubMed
Scolding, N. J., Morgan, B. P., and Compston, D. A. (1998b). The expression of complement regulatory proteins by adult human oligodendrocytes. J Neuroimmunol 84, 69–75.CrossRefGoogle ScholarPubMed
Scolding, N. J., Morgan, B. P., Houston, W. A. J., Linington, C., Campbell, A. K., and Compston, D. A. S. (1989). Vesicular removal by oligodendrocytes of membrane attack complexes formed by activated complement. Nature 339, 620–622.CrossRefGoogle ScholarPubMed
Scolding, N. J., Rayner, P. J., and Compston, D. A. (1999). Identification of A2B5-positive putative oligodendrocyte progenitor cells and A2B5-positive astrocytes in adult human white matter. Neuroscience 89, 1–4.CrossRefGoogle ScholarPubMed
Scolding, N. J., Rayner, P. J., Sussman, J., Shaw, C., and Compston, D. A. S. (1995). A proliferative adult human oligodendrocyte progenitor. NeuroReport 6, 441–445.CrossRefGoogle ScholarPubMed
Scott, G. S., Spitsin, S. V., Kean, R. B., Mikheeva, T., Koprowski, H., and Hooper, D. C. (2002). Therapeutic intervention in experimental allergic encephalomyelitis by administration of uric acid precursors. Proc Natl Acad Sci USA 99, 16303–16308.CrossRefGoogle ScholarPubMed
Scott, G. S., Virag, L., Szabo, C., and Hooper, D. C. (2003). Peroxynitrite-induced oligodendrocyte toxicity is not dependent on poly(ADP-ribose) polymerase activation. Glia 41, 105–116.CrossRefGoogle Scholar
Seeman, P., Mazanec, R., Ctvrteckova, M., and Smilkova, D. (2001). Charcot-Marie-Tooth type X: a novel mutation in the Cx32 gene with central conduction slowing. Int J Mol Med 8, 461–468.Google ScholarPubMed
Segal, D., Koschnick, J. R., Slegers, L. H., and Hof, P. R. (2007). Oligodendrocyte pathophysiology: a new view of schizophrenia. Int J Neuropsychopharmacol 10, 503–511.CrossRefGoogle ScholarPubMed
Segovia, K. N., McClure, M., Moravec, M., et al. (2008). Arrested oligodendrocyte lineage maturation in chronic perinatal white matter injury. Ann Neurol 63, 520–530.CrossRefGoogle ScholarPubMed
Seilheimer, B. and Schachner, M. (1988). Studies of adhesion molecules mediating interactions between cells of peripheral nervous system indicate a major role for L1 in mediating sensory neuron growth on Schwann cells in culture. J Cell Biol 107, 341–351.CrossRefGoogle Scholar
Seilheimer, B., Persohn, E., and Schachner, M. (1989). Antibodies to the L1 adhesion molecule inhibit Schwann cell ensheathment of neurons in vitro. J Cell Biol 109, 3095–3103.CrossRefGoogle ScholarPubMed
Seki, Y., Feustel, P. J., Keller, R. W., Tranmer, B. I., and Kimelberg, H. K. (1999). Inhibition of ischemia-induced glutamate release in rat striatum by dihydrokinate and an anion channel blocker. Stroke 30, 433–440.CrossRefGoogle ScholarPubMed
Sellers, D. L., Maris, D. O., and Horner, P. J. (2009). Postinjury niches induce temporal shifts in progenitor fates to direct lesion repair after spinal cord injury. J Neurosci 29, 6722–6733.CrossRefGoogle ScholarPubMed
Selmaj, K. W. (2000). Tumour necrosis factor and anti-tumour necrosis factor approach to inflammatory demyelinating diseases of the central nervous system. Ann Rheum Dis 59 Suppl 1, i94–102.CrossRefGoogle ScholarPubMed
Selmaj, K., Brosnan, C. F., and Raine, C. S. (1991a). Colocalization of lymphocytes bearing gamma/delta T-cell receptor and heat shock protein hsp 65+ oligodendrocytes in multiple sclerosis. Proc Natl Acad Sci USA 88, 6452–6456.CrossRefGoogle Scholar
Selmaj, K., Raine, C. S., and Cross, A. H. (1991b). Anti-tumor necrosis factor therapy abrogates autoimmune demyelination. Ann Neurol 30, 694–700.CrossRefGoogle ScholarPubMed
Shao, Z., Browning, J. L., Lee, X., et al. (2005). TAJ/TROY, an orphan TNF receptor family member, binds Nogo-66 receptor 1 and regulates axonal regeneration. Neuron 45, 353–359.CrossRefGoogle ScholarPubMed
Sharp, F. R. and Bernaudin, M. (2004). HIF1 and oxygen sensing in the brain. Nat Rev Neurosci 5, 437–448.CrossRefGoogle Scholar
Sheikh, K. A., Sun, J., Liu, Y., et al. (1999). Mice lacking complex gangliosides develop Wallerian degeneration and myelination defects. Proc Natl Acad Sci USA 96, 7532–7537.CrossRefGoogle ScholarPubMed
Sherman, D. L. and Brophy, P. J. (2005). Mechanisms of axon ensheathment and myelin growth. Nat Rev Neurosci 6, 683–690.CrossRefGoogle ScholarPubMed
Sherman, D. L., Tait, S., Melrose, S., et al. (2005). Neurofascins are required to establish axonal domains for saltatory conduction. Neuron 48, 737–742.CrossRefGoogle ScholarPubMed
Shi, Y., Do, J. T., Desponts, C., Hahm, H. S., Scholer, H. R., and Ding, S. (2008). A combined chemical and genetic approach for the generation of induced pluripotent stem cells. Cell Stem Cell 2, 525–528.CrossRefGoogle ScholarPubMed
Siegel, G. J., Agranoff, B. W., Fisher, S. K., Albers, R. W., and Uhler, M. D. (1999). Basic Neurochemistry: Molecular, Cellular and Medical Aspects. (Philadelphia, PA: Lippincott–Raven Publishers.)Google Scholar
Silver, J. and Miller, J. H. (2004). Regeneration beyond the glial scar. Nat Rev Neurosci 5, 146–156.CrossRefGoogle ScholarPubMed
Silverstein, F. S., Naik, B., and Simpson, J. (1991). Hypoxia-ischemia stimulates hippocampal glutamate efflux in perinatal rat brain: an in vivo microdialysis study. Pediatr Res 30, 587–590.CrossRefGoogle Scholar
Simard, A. R. and Rivest, S. (2006). Neuroprotective properties of the innate immune system and bone marrow stem cells in Alzheimer's disease. Mol Psychiatry 11, 327–335.CrossRefGoogle ScholarPubMed
Simonen, M., Pedersen, V., Weinmann, O., et al. (2003). Systemic deletion of the myelin-associated outgrowth inhibitor Nogo-A improves regenerative and plastic responses after spinal cord injury. Neuron 38, 201–211.CrossRefGoogle ScholarPubMed
Simons, M. and Trajkovic, K. (2006). Neuron-glia communication in the control of oligodendrocyte function and myelin biogenesis. J Cell Sci 119, 4381–4389.CrossRefGoogle ScholarPubMed
Simons, M. and Trotter, J. (2007). Wrapping it up: the cell biology of myelination. Curr Opin Neurobiol 17, 533–540.CrossRefGoogle ScholarPubMed
Simpson, P. B. and Armstrong, R. C. (1999). Intracellular signals and cytoskeletal elements involved in oligodendrocyte progenitor migration. Glia 26, 22–35.3.0.CO;2-M>CrossRefGoogle ScholarPubMed
Singec, I. and Snyder, E. Y. (2008). Inflammation as a matchmaker: revisiting cell fusion. Nat Cell Biol 10, 503–505.CrossRefGoogle ScholarPubMed
Singh, A. K., Wilson, M. T., Hong, S., et al. (2001). Natural killer T cell activation protects mice against experimental autoimmune encephalomyelitis. J Exp Med 194, 1801–1811.CrossRefGoogle ScholarPubMed
Sinibaldi, L., Luca, A., Bellacchio, E., et al. (2004). Mutations of the Nogo-66 receptor (RTN4R) gene in schizophrenia. Hum Mutat 24, 534–535.CrossRefGoogle Scholar
Sloane, J. A. and Vartanian, T. K. (2007a). Myosin Va controls oligodendrocyte morphogenesis and myelination. J Neurosci 27, 11366–11375.CrossRefGoogle ScholarPubMed
Sloane, J. A. and Vartanian, T. K. (2007b). WAVE1 and regulation of actin nucleation in myelination. Neuroscientist 13, 486–491.CrossRefGoogle ScholarPubMed
Sloane, J. A., Hinman, J. D., Lubonia, M., Hollander, W., and Abraham, C. R. (2003). Age-dependent myelin degeneration and proteolysis of oligodendrocyte proteins is associated with the activation of calpain-1 in the rhesus monkey. J Neurochem 84, 157–168.CrossRefGoogle ScholarPubMed
Smirnova, L., Grafe, A., Seiler, A., Schumacher, S., Nitsch, R., and Wulczyn, F. G. (2005). Regulation of miRNA expression during neural cell specification. Eur J Neurosci 21, 1469–1477.CrossRefGoogle ScholarPubMed
Smith, K. J., Blakemore, W. F., and McDonald, W. I. (1979). Central remyelination restores secure conduction. Nature 280, 395–396.CrossRefGoogle ScholarPubMed
Smith, K. J., Blakemore, W. F., and McDonald, W. I. (1981). The restoration of conduction by central remyelination. Brain 104, 383–404.CrossRefGoogle ScholarPubMed
Smith, K. J., Bostock, H., and Hall, S. M. (1982). Saltatory conduction precedes remyelination in axons demyelinated with lysophosphatidyl choline. J Neurol Sci 54, 13–31.CrossRefGoogle ScholarPubMed
Smith, T., Groom, A., Zhu, B., and Turski, L. (2000). Autoimmune encephalomyelitis ameliorated by AMPA antagonists. Nat Med 6, 62–66.CrossRefGoogle ScholarPubMed
Snethen, H., Love, S., and Scolding, N. (2008). Disease-responsive neural precursor cells are present in multiple sclerosis lesions. Regenerative Med 3, 835–847.CrossRefGoogle ScholarPubMed
Soane, L., Cho, H. J., Niculescu, F., Rus, H., and Shin, M. L. (2001). C5b-9 terminal complement complex protects oligodendrocytes from death by regulating Bad through phosphatidylinositol 3-kinase/Akt pathway. J Immunol 167, 2305–2311.CrossRefGoogle ScholarPubMed
Sohn, J., Natale, J., Chew, L. J., et al. (2006). Identification of Sox17 as a transcription factor that regulates oligodendrocyte development. J Neurosci 26, 9722–9735.CrossRefGoogle ScholarPubMed
Sokolov, B. P. (2007). Oligodendroglial abnormalities in schizophrenia, mood disorders and substance abuse. Comorbidity, shared traits, or molecular phenocopies? Int J Neuropsychopharmacol 10, 547–555.CrossRefGoogle ScholarPubMed
Soldan, M. M., Warrington, A. E., Bieber, A. J., et al. (2003). Remyelination-promoting antibodies activate distinct Ca2+ influx pathways in astrocytes and oligodendrocytes: relationship to the mechanism of myelin repair. Mol Cell Neurosci 22, 14–24.CrossRefGoogle Scholar
Somjen, G. G. (1988). Nervenkitt: notes on the history of the concept of neuroglia. Glia 1, 2–9.CrossRefGoogle ScholarPubMed
Song, J., Goetz, B. D., Baas, P. W., and Duncan, I. D. (2001). Cytoskeletal reorganization during the formation of oligodendrocyte processes and branches. Mol Cell Neurosci 17, 624–636.CrossRefGoogle ScholarPubMed
Song, J., O'Connor, L. T., Yu, W., Baas, P. W., and Duncan, I. D. (1999). Microtubule alterations in cultured taiep rat oligodendrocytes lead to deficits in myelin membrane formation. J Neurocytol 28, 671–683.CrossRefGoogle ScholarPubMed
Song, S. K., Sun, S. W., Ramsbottom, M. J., Chang, C., Russell, J., and Cross, A. H. (2002). Dysmyelination revealed through MRI as increased radial (but unchanged axial) diffusion of water. Neuroimage 17, 1429–1436.CrossRefGoogle Scholar
Song, S. K., Yoshino, J., Le, T. Q., et al. (2005). Demyelination increases radial diffusivity in corpus callosum of mouse brain. Neuroimage 26, 132–140.CrossRefGoogle ScholarPubMed
Sontheimer, H. (1991). Astrocytes, as well as neurons, express a diversity of ion channels. Can J Physiol Pharmacol 70, s223–s238.CrossRefGoogle Scholar
Southwood, C. M., Peppi, M., Dryden, S., Tainsky, M. A., and Gow, A. (2007). Microtubule deacetylases, SirT2 and HDAC6, in the nervous system. Neurochem Res 32, 187–195.CrossRefGoogle Scholar
Spassky, N., Goujet-Zalc, C., Parmantier, E., et al. (1998). Multiple restricted origin of oligodendrocytes. J Neurosci 18, 8331–8343.CrossRefGoogle ScholarPubMed
Spiegel, I. and Peles, E. (2002). Cellular junctions of myelinated nerves (Review). Mol Membr Biol 19, 95–101.CrossRefGoogle Scholar
Spiegel, I., Adamsky, K., Eshed, Y., et al. (2007). A central role for Necl4 (SynCAM4) in Schwann cell-axon interaction and myelination. Nat Neurosci 10, 861–869.CrossRefGoogle ScholarPubMed
Spillantini, M. G., Goedert, M., Crowther, R. A., Murrell, J. R., Farlow, M. R., and Ghetti, B. (1997). Familial multiple system tauopathy with presenile dementia: a disease with abundant neuronal and glial tau filaments. Proc Natl Acad Sci USA 94, 4113–4118.CrossRefGoogle ScholarPubMed
Spillmann, A. A., Bandtlow, C. E., Lottspeich, F., Keller, F., and Schwab, M. E. (1998). Identification and characterization of a bovine neurite growth inhibitor (bNI-220). J Biol Chem 273, 19283–19293.CrossRefGoogle Scholar
Squier, T. C. (2001). Oxidative stress and protein aggregation during biological aging. Exp Gerontol 36, 1539–1550.CrossRefGoogle ScholarPubMed
Srinivasan, R., Sailasuta, N., Hurd, R., Nelson, S., and Pelletier, D. (2005). Evidence of elevated glutamate in multiple sclerosis using magnetic resonance spectroscopy at 3 T. Brain 128, 1016–1025.CrossRefGoogle ScholarPubMed
Stadelmann, C., Ludwin, S., Tabira, T., et al. (2005). Tissue preconditioning may explain concentric lesions in Balo's type of multiple sclerosis. Brain 128, 979–987.CrossRefGoogle ScholarPubMed
Stefansson, H., Sigurdsson, E., Steinthorsdottir, V., et al. (2002). Neuregulin 1 and susceptibility to schizophrenia. Am J Hum Genet 71, 877–892.CrossRefGoogle Scholar
Steinman, L. (1996). A few autoreactive cells in an autoimmune infiltrate control a vast population of nonspecific cells: a tale of smart bombs and the infantry. Proc Natl Acad Sci USA 93, 2253–2256.CrossRefGoogle Scholar
Sternberger, N. H., Itoyama, Y., Koco, M. W., and Webster, H. D. (1978). Myelin basic protein demonstrated immunocytochemically in oligodendroglia prior to myelin sheath formation. Proc Natl Acad Sci USA 75, 2521–2524.CrossRefGoogle ScholarPubMed
Sternberger, N. H., Quarles, R. H., Itoyama, Y., and Webster, H. D. (1979). Myelin-associated glycoprotein demonstrated immunocytochemically in myelin and myelin-forming cells of developing rats. Proc Natl Acad Sci USA 76, 1510–1514.CrossRefGoogle Scholar
Stevens, B. and Fields, R. D. (2000). Response of Schwann cells to action potentials in development. Science 287, 2267–2271.CrossRefGoogle ScholarPubMed
Stevens, B., Porta, S., Haak, L. L., Gallo, V., and Fields, R. D. (2002). Adenosine: a neuron-glial transmitter promoting myelination in the CNS in response to action potentials. Neuron 36, 855–868.CrossRefGoogle ScholarPubMed
Stidworthy, M. F., Genoud, S., Suter, U., Mantei, N., and Franklin, R. J. (2003). Quantifying the early stages of remyelination following cuprizone-induced demyelination. Brain Pathol 13, 329–339.CrossRefGoogle ScholarPubMed
Stoffel, W., Boison, D., and Bussow, H. (1997). Functional analysis in vivo of the double mutant mouse deficient in both proteolipid protein (PLP) and myelin basic protein (MBP) in the central nervous system. Cell Tissue Res 289, 195–206.CrossRefGoogle Scholar
Stokes, B. T. and Jakeman, L. B. (2002). Experimental modelling of human spinal cord injury: a model that crosses the species barrier and mimics the spectrum of human cytopathology. Spinal Cord 40, 101–109.CrossRefGoogle ScholarPubMed
Stokes, B. T., Garwood, M., and Walters, P. (1981). Oxygen fields in specific spinal loci of the canine spinal cord. Am J Physiol 240, H761–H766.Google ScholarPubMed
Stoll, G. and Muller, H. W. (1999). Nerve injury, axonal degeneration and neural regeneration: basic insights. Brain Pathol 9, 313–325.CrossRefGoogle ScholarPubMed
Stoll, G., Schroeter, M., Jander, S., et al. (2004). Lesion-associated expression of transforming growth factor-beta-2 in the rat nervous system: evidence for down-regulating the phagocytic activity of microglia and macrophages. Brain Pathol 14, 51–58.CrossRefGoogle ScholarPubMed
Stoll, G., Trapp, B. D., and Griffin, J. W. (1989). Macrophage function during Wallerian degeneration of rat optic nerve: clearance of degenerating myelin and Ia expression. J Neurosci 9, 2327–2335.CrossRefGoogle ScholarPubMed
Storck, T., Schulte, S., Hofmann, K., and Stoffel, W. (1992). Structure, expression, and functional analysis of a Na+- dependent glutamate/aspartate transporter from rat brain. Proc Natl Acad Sci USA 89, 10955–10959.CrossRefGoogle ScholarPubMed
Stys, P. K. (2004). White matter injury mechanisms. Curr Mol Med 4, 113–130.CrossRefGoogle ScholarPubMed
Sugai, T., Kawamura, M., Iritani, S., et al. (2004). Prefrontal abnormality of schizophrenia revealed by DNA microarray: impact on glial and neurotrophic gene expression. Ann N Y Acad Sci 1025, 84–91.CrossRefGoogle ScholarPubMed
Sulaiman, O. A. R., Boyd, J. G., and Gordon, T. (2005). Axonal regeneration in the peripheral nervous system of mammals. In Neuroglia, Kettenmann, H. and Ransom, B. R., eds. (New York: Oxford University Press), pp. 454–466.Google Scholar
Sullivan, C. D. and Geisert, E. E. (1998). Expression of rat target of the antiproliferative antibody (TAPA) in the developing brain. J Comp Neurol 396, 366–380.3.0.CO;2-0>CrossRefGoogle Scholar
Suzuki, K. (2003). Globoid cell leukodystrophy (Krabbe's disease): update. J Child Neurol 18, 595–603.CrossRefGoogle ScholarPubMed
Svaren, J. and Meijer, D. (2008). The molecular machinery of myelin gene transcription in Schwann cells. Glia 56, 1541–1551.CrossRefGoogle ScholarPubMed
Swanson, R. A., Miller, J. W., Rothstein, J. D., Farrell, K., Stein, B. A., and Longuemare, M. C. (1997). Neuronal regulation of glutamate transporter subtype expression in astrocytes. J Neurosci 17, 932–940.CrossRefGoogle ScholarPubMed
Szatkowski, M., Barbour, B., and Attwell, D. (1990). Non-vesicular release of glutamate from glial cells by reversed electrogenic glutamate uptake. Nature 348, 443–446.CrossRefGoogle ScholarPubMed
Tabira, T., Cullen, M. J., Reier, P. J., and Webster, H. D. (1978). An experimental analysis of interlamellar tight junctions in amphibian and mammalian CNS myelin. J Neurocytol 7, 489–503.CrossRefGoogle Scholar
Tait, S., Gunn-Moore, F., Collinson, J. M., et al. (2000). An oligodendroctye cell adhesion molecule at the site of assembly of the paranodal axo-glial junction. J Cell Biol 150, 657–666.CrossRefGoogle Scholar
Takahashi, K., Tanabe, K., Ohnuki, M., et al. (2007). Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 131, 861–872.CrossRefGoogle ScholarPubMed
Takano, R., Misu, T., Takahashi, T., Izumiyama, M., Fujihara, K., and Itoyama, Y. (2008). A prominent elevation of glial fibrillary acidic protein in the cerebrospinal fluid during relapse in neuromyelitis optica. Tohoku J Exp Med 215, 55–59.CrossRefGoogle ScholarPubMed
Takebayashi, H., Nabeshima, Y., Yoshida, S., Chisaka, O., and Ikenaka, K. (2002). The basic helix-loop-helix factor olig2 is essential for the development of motoneuron and oligodendrocyte lineages. Curr Biol 12, 1157–1163.CrossRefGoogle ScholarPubMed
Takeda, Y., Asou, H., Murakami, Y., Miura, M., Kobayashi, M., and Uyemura, K. (1996). A nonneuronal isoform of cell adhesion molecule L1: tissue-specific expression and functional analysis. J Neurochem 66, 2338–2349.CrossRefGoogle ScholarPubMed
Talbott, J. F., Loy, D. N., Liu, Y., et al. (2005). Endogenous Nkx2.2+/Olig2+ oligodendrocyte precursor cells fail to remyelinate the demyelinated adult rat spinal cord in the absence of astrocytes. Exp Neurol 192, 11–24.CrossRefGoogle ScholarPubMed
Talos, D. M., Fishman, R. E., Park, H., et al. (2006). Developmental regulation of alpha-amino-3-hydroxy-5-methyl-4-isoxazole-propionic acid receptor subunit expression in forebrain and relationship to regional susceptibility to hypoxic/ischemic injury. I. Rodent cerebral white matter and cortex. J Comp Neurol 497, 42–60.CrossRefGoogle ScholarPubMed
Tambuyzer, B. R., Ponsaerts, P., and Nouwen, E. J. (2009). Microglia: gatekeepers of central nervous system immunology. J Leukoc Biol 85, 352–370.CrossRefGoogle ScholarPubMed
Tanaka, J. and Sobue, K. (1994). Localization and characterization of gelsolin in nervous tissues: gelsolin is specifically enriched in myelin-forming cells. J Neurosci 14, 1038–1052.CrossRefGoogle ScholarPubMed
Tanaka, K., Watase, K., Manabe, T., et al. (1997). Epilepsy and exacerbation of brain injury in mice lacking the glutamate transporter GLT-1. Science 276, 1699–1702.CrossRefGoogle ScholarPubMed
Targett, M., Sussman, J., Scolding, N., O'Leary, M. T., Compston, D. A., and Blakemore, W. F. (1996). Failure to achieve remyelination of demyelinated rat axons following transplantation of glial cells obtained from the adult human brain. Neuropathol Appl Neurobiol 22, 199–206.CrossRefGoogle ScholarPubMed
Tator, C. H. (1995). Update on the pathophysiology and pathology of acute spinal cord injury. Brain Pathol 5, 407–413.CrossRefGoogle ScholarPubMed
Taveggia, C., Thaker, P., Petrylak, A., et al. (2008). Type III neuregulin-1 promotes oligodendrocyte myelination. Glia 56, 284–293.CrossRefGoogle ScholarPubMed
Taveggia, C., Zanazzi, G., Petrylak, A., et al. (2005). Neuregulin-1 type III determines the ensheathment fate of axons. Neuron 47, 681–694.CrossRefGoogle ScholarPubMed
Tekkök, S. B. and Goldberg, M. P. (2001). AMPA/Kainate receptor activation mediates hypoxic oligodendrocyte death and axonal injury in cerebral white matter. J Neurosci 21, 4237–4248.CrossRefGoogle ScholarPubMed
Temple, S. and Raff, M. C. (1986). Clonal analysis of oligodendrocyte development in culture: evidence for a developmental clock that counts cell divisions. Cell 44, 773–779.CrossRefGoogle ScholarPubMed
Teng, F. Y. and Tang, B. L. (2005). Why do Nogo/Nogo-66 receptor gene knockouts result in inferior regeneration compared to treatment with neutralizing agents? J Neurochem 94, 865–874.CrossRefGoogle ScholarPubMed
Terada, N., Baracskay, K., Kinter, M., et al. (2002a). The tetraspanin protein, CD9, is expressed by progenitor cells committed to oligodendrogenesis and is linked to beta1 integrin, CD81, and Tspan-2. Glia 40, 350–359.CrossRefGoogle ScholarPubMed
Terada, N., Hamazaki, T., Oka, M., et al. (2002b). Bone marrow cells adopt the phenotype of other cells by spontaneous cell fusion. Nature 416, 542–545.CrossRefGoogle ScholarPubMed
Terada, N., Kidd, G. J., Kinter, M., Bjartmar, C., Moran-Jones, K., and Trapp, B. D. (2005). Beta(IV) tubulin is selectively expressed by oligodendrocytes in the central nervous system. Glia 50, 212–222.CrossRefGoogle ScholarPubMed
Thomson, J. M., Parker, J., Perou, C. M., and Hammond, S. M. (2004). A custom microarray platform for analysis of microRNA gene expression. Nat Methods 1, 47–53.CrossRefGoogle ScholarPubMed
Thorburne, S. K. and Juurlink, B. J. H. (1996). Low gluthathione and high iron govern the susceptibility of oligodendroglial precursosrs to oxidative stress. J Neuroimmunol 67, 1014–1022.Google Scholar
Timsit, S., Martinez, S., Allinquant, B., Peyron, F., Puelles, L., and Zalc, B. (1995). Oligodendrocytes originate in a restricted zone of the embryonic ventral neural tube defined by DM-20 mRNA expression. J Neurosci 15, 1012–1024.CrossRefGoogle Scholar
Timsit, S., Sinoway, M. P., Levy, L., et al. (1992). The DM20 protein of myelin: intracellular and surface expression patterns in transfectants. J Neurochem 58, 1936–1942.CrossRefGoogle ScholarPubMed
Ting, A. E., Mays, R. W., Frey, M. R., Hof, W. V., Medicetty, S., and Deans, R. (2008). Therapeutic pathways of adult stem cell repair. Crit Rev Oncol Hematol 65, 81–93.CrossRefGoogle ScholarPubMed
Tkachev, D., Mimmack, M. L., Ryan, M. M., et al. (2003). Oligodendrocyte dysfunction in schizophrenia and bipolar disorder. Lancet 362, 798–805.CrossRefGoogle ScholarPubMed
Tofaris, G. K., Patterson, P. H., Jessen, K. R., and Mirsky, R. (2002). Denervated Schwann cells attract macrophages by secretion of leukemia inhibitory factor (LIF) and monocyte chemoattractant protein-1 in a process regulated by interleukin-6 and LIF. J Neurosci 22, 6696–6703.CrossRefGoogle Scholar
Tom, V. J., Steinmetz, M. P., Miller, J. H., Doller, C. M., and Silver, J. (2004). Studies on the development and behavior of the dystrophic growth cone, the hallmark of regeneration failure, in an in vitro model of the glial scar and after spinal cord injury. J Neurosci 24, 6531–6539.CrossRefGoogle Scholar
Tomonari, A., Tojo, A., Adachi, D., et al. (2003). Acute disseminated encephalomyelitis (ADEM) after allogeneic bone marrow transplantation for acute myeloid leukemia. Ann Hematol 82, 37–40.Google ScholarPubMed
Torcia, M., Bracci-Laudiero, L., Lucibello, M., et al. (1996). Nerve growth factor is an autocrine survival factor for memory B lymphocytes. Cell 85, 345–356.CrossRefGoogle ScholarPubMed
Tornatore, C., Berger, J. R., Houff, S. A., et al. (1992). Detection of JC virus DNA in peripheral lymphocytes from patients with and without progressive multifocal leukoencephalopathy. Ann Neurol 31, 454–462.CrossRefGoogle ScholarPubMed
Totoiu, M. O. and Keirstead, H. S. (2005). Spinal cord injury is accompanied by chronic progressive demyelination. J Comp Neurol 486, 373–383.CrossRefGoogle ScholarPubMed
Traka, M., Dupree, J. L., Popko, B., and Karagogeos, D. (2002). The neuronal adhesion protein TAG-1 is expressed by Schwann cells and oligodendrocytes and is localized to the juxtaparanodal region of myelinated fibers. J Neurosci 22, 3016–3024.CrossRefGoogle ScholarPubMed
Traka, M., Goutebroze, L., Denisenko, N., et al. (2003). Association of TAG-1 with Caspr2 is essential for the molecular organization of juxtaparanodal regions of myelinated fibers. J Cell Biol 162, 1161–1172.CrossRefGoogle ScholarPubMed
Tran, P. B. and Miller, R. J. (2003). Chemokine receptors: signposts to brain development and disease. Nat Rev Neurosci 4, 444–455.CrossRefGoogle ScholarPubMed
Trapp, B. and Kidd, G. (2004). Structure of the myelinated axon. In Myelin Biology and Disorders, Volume 1, Lazzarini, R. A., Griffin, J. W., Lassmann, H., Nave, K. A., Miller, R., and Trapp, B., eds. (San Diego, CA: Academic Press), pp. 3–27.Google Scholar
Trapp, B. D. and Nave, K. A. (2008). Multiple sclerosis: an immune or neurodegenerative disorder? Annu Rev Neurosci 31, 247–269.CrossRefGoogle ScholarPubMed
Trapp, B. D. and Stys, P. K. (2009). Virtual hypoxia and chronic necrosis of demyelinated axons in multiple sclerosis. Lancet Neurol 8, 280–291.CrossRefGoogle ScholarPubMed
Trapp, B. D., Andrews, S. B., Cootauco, C., and Quarles, R. H. (1989). The myelin-associated glycoprotein is enriched in multivesicular bodies and periaxonal membranes of actively myelinating oligodendrocytes. J Cell Biol 109, 2417–2426.CrossRefGoogle ScholarPubMed
Trapp, B. D., Bernier, L., Andrews, S. B., and Colman, D. R. (1988). Cellular and subcellular distribution of 2′,3′ cyclic nucleotide 3′ phosphodiesterase and its mRNA in the rat nervous system. J Neurochem 51, 859–868.CrossRefGoogle Scholar
Trapp, B. D., Itoyama, Y., MacIntosh, T. D., and Quarles, R. H. (1983). P2 protein in oligodendrocytes and myelin of the rabbit central nervous system. J Neurochem 40, 47–54.CrossRefGoogle ScholarPubMed
Trapp, B. D., Moench, T., Pulley, M., Barbosa, E., Tennekoon, G., and Griffin, J. W. (1987). Spatial segregation of mRNA encoding myelin-specific proteins. Proc Natl Acad Sci USA 84, 7773–7777.CrossRefGoogle ScholarPubMed
Trapp, B. D., Nishiyama, A., Cheng, D., and Macklin, W. (1997). Differentiation and death of premyelinating oligodendrocytes in developing rodent brain. J Cell Biol 137, 459–468.CrossRefGoogle ScholarPubMed
Trapp, B. D., Peterson, J., Ransohoff, R. M., Rudick, R. A., Mork, S., and Bo, L. (1998). Axon transection in the lesions of multiple sclerosis. N Engl J Med 338, 278–285.CrossRefGoogle ScholarPubMed
Trebst, C., Heine, S., Lienenklaus, S., et al. (2007). Lack of interferon-beta leads to accelerated remyelination in a toxic model of central nervous system demyelination. Acta Neuropathol 114, 587–596.CrossRefGoogle Scholar
Trotter, J. (2005). NG2-positive cells in CNS function and the pathological role of antibodies against NG2 in demyelinating diseases. J Neurol Sci 233, 37–42.CrossRefGoogle ScholarPubMed
Trotter, J., Bitter-Suermann, D., and Schachner, M. (1989). Differentiation-regulated loss of the polysialylated embryonic form and expression of the different polypeptides of the neural cell adhesion molecule by cultured oligodendrocytes and myelin. J Neurosci Res 22, 369–383.CrossRefGoogle ScholarPubMed
Trotter, J. L., Collins, K. G., and Veen, R. (1991). Serum cytokine levels in chronic progressive multiple sclerosis: interleukin-2 levels parallel tumor necrosis factor-α levels. J Neuroimmunol 33, 29–36.CrossRefGoogle ScholarPubMed
Tsai, H., Macklin, W. B., and Miller, R. H. (2006). Netrin 1 is required for the normal development of spinal cord oligodendrocytes. J Neurosci 26, 1913–1922.CrossRefGoogle ScholarPubMed
Tsai, H. H., Frost, E., To, V., et al. (2002). The chemokine receptor CXCR2 controls positioning of oligodendrocyte precursors in developing spinal cord by arresting their migration. Cell 110, 373–383.CrossRefGoogle ScholarPubMed
Tsai, H. H., Macklin, W. B., and Miller, R. H. (2009). Distinct modes of migration position oligodendrocyte precursors for localized cell division in the developing spinal cord. J Neurosci Res. 87, 3320–3330.CrossRefGoogle ScholarPubMed
Tsai, H. H., Tessier-Lavigne, M., and Miller, R. H. (2003). Netrin 1 mediates spinal cord oligodendrocyte precursor dispersal. Development 130, 2095–2105.CrossRefGoogle ScholarPubMed
Tsuru, T., Mizuguchi, M., Ohkubo, Y., Itonaga, N., and Momoi, M. Y. (2000). Acute disseminated encephalomyelitis after live rubella vaccination. Brain Dev 22, 259–261.CrossRefGoogle ScholarPubMed
Tzingounis, A. V. and Wadiche, J. I. (2007). Glutamate transporters: confining runaway excitation by shaping synaptic transmission. Nat Rev Neurosci 8, 935–947.CrossRefGoogle ScholarPubMed
Ueda, H., Levine, J. M., Miller, R. H., and Trapp, B. D. (1999). Rat optic nerve oligodendrocytes develop in the absence of viable retinal ganglion cell axons. J Cell Biol 146, 1365–1374.CrossRefGoogle ScholarPubMed
Uhlenberg, B., Schuelke, M., Ruschendorf, F., et al. (2004). Mutations in the gene encoding gap junction protein alpha 12 (connexin 46.6) cause Pelizaeus-Merzbacher-like disease. Am J Hum Genet 75, 251–260.CrossRefGoogle ScholarPubMed
Ulzheimer, J. C., Peles, E., Levinson, S. R., and Martini, R. (2004). Altered expression of ion channel isoforms at the node of Ranvier in P0-deficient myelin mutants. Mol Cell Neurosci 25, 83–94.CrossRefGoogle ScholarPubMed
Uschkureit, T., Sporkel, O., Stracke, J., Bussow, H., and Stoffel, W. (2000). Early onset of axonal degeneration in double (plp–/–mag–/–) and hypomyelinosis in triple (plp–/–mbp–/–mag–/–) mutant mice. J Neurosci 20, 5225–5233.CrossRefGoogle ScholarPubMed
Valerio, A., Ferrario, M., Dreano, M., Garotta, G., Spano, P., and Pizzi, M. (2002). Soluble interleukin-6 (IL-6) receptor/IL-6 fusion protein enhances in vitro differentiation of purified rat oligodendroglial lineage cells. Mol Cell Neurosci 21, 602–615.CrossRefGoogle ScholarPubMed
Vallstedt, A., Klos, J. M., and Ericson, J. (2005). Multiple dorsoventral origins of oligodendrocyte generation in the spinal cord and hindbrain. Neuron 45, 55–67.CrossRefGoogle ScholarPubMed
Valk, P. and Groot, C. J. (2000). Staging of multiple sclerosis (MS) lesions: pathology of the time frame of MS. Neuropathol Appl Neurobiol 26, 2–10.CrossRefGoogle Scholar
Vanderlugt, C. L. and Miller, S. D. (2002). Epitope spreading in immune-mediated diseases: implications for immunotherapy. Nat Rev Immunol 2, 85–95.CrossRefGoogle ScholarPubMed
Vanguri, P., Koski, C. L., Silverman, B., and Shin, M. L. (1982). Complement activation by isolated myelin: activation of the classical pathway in the absence of myelin-specific antibodies. Proc Natl Acad Sci USA 79, 3290–3294.CrossRefGoogle ScholarPubMed
Vargas, M. E. and Barres, B. A. (2007). Why is Wallerian degeneration in the CNS so slow? Annu Rev Neurosci 30, 153–179.CrossRefGoogle ScholarPubMed
Vartanian, T., Fischbach, G., and Miller, R. (1999). Failure of spinal cord oligodendrocyte development in mice lacking neuregulin. Proc Natl Acad Sci USA 96, 731–735.CrossRefGoogle ScholarPubMed
Vartanian, T., Goodearl, A., Viehover, A., and Fischbach, G. (1997). Axonal neuregulin signal cells of the oligodendrocyte lineage through activation of HER4 and Schwann cells through HER2 and HER3. J Cell Biol 137, 211–220.CrossRefGoogle ScholarPubMed
Vassilopoulos, G., Wang, P. R., and Russell, D. W. (2003). Transplanted bone marrow regenerates liver by cell fusion. Nature 422, 901–904.CrossRefGoogle ScholarPubMed
Vaughn, J. E. and Pease, D. C. (1970). Electron microscopic studies of wallerian degeneration in rat optic nerves. II. Astrocytes, oligodendrocytes and adventitial cells. J Comp Neurol 140, 207–226.CrossRefGoogle ScholarPubMed
Vela, J. M., Molina-Holgado, E., Arevalo-Martin, A., Almazan, G., and Guaza, C. (2002). Interleukin-1 regulates proliferation and differentiation of oligodendrocyte progenitor cells. Mol Cell Neurosci 20, 489–502.CrossRefGoogle ScholarPubMed
Venkatesh, K., Chivatakarn, O., Lee, H., et al. (2005). The Nogo-66 receptor homolog NgR2 is a sialic acid-dependent receptor selective for myelin-associated glycoprotein. J Neurosci 25, 808–822.CrossRefGoogle ScholarPubMed
Vergelli, M., Le, H., Noort, J. M., Dhib-Jalbut, S., McFarland, H., and Martin, R. (1996). A novel population of CD4+CD56+ myelin-reactive T cells lyses target cells expressing CD56/neural cell adhesion molecule. J Immunol 157, 679–688.Google ScholarPubMed
Villoslada, P., Hauser, S. L., Bartke, I., et al. (2000). Human nerve growth factor protects common marmosets against autoimmune encephalomyelitis by switching the balance of T helper cell type 1 and 2 cytokines within the central nervous system. J Exp Med 191, 1799–1806.CrossRefGoogle ScholarPubMed
Vincent, T., Saikali, P., Cayrol, R., et al. (2008). Functional consequences of neuromyelitis optica-IgG astrocyte interactions on blood-brain barrier permeability and granulocyte recruitment. J Immunol 181, 5730–5737.CrossRefGoogle ScholarPubMed
Vitkovic, L., Konsman, J. P., Bockaert, J., Dantzer, R., Homburger, V., and Jacque, C. (2000). Cytokine signals propagate through the brain. Mol Psychiatry 5, 604–615.CrossRefGoogle Scholar
Vogler, S., Pahnke, J., Rousset, S., et al. (2006). Uncoupling protein 2 has protective function during experimental autoimmune encephalomyelitis. Am J Pathol 168, 1570–1575.CrossRefGoogle ScholarPubMed
Volpe, J. J. (2003). Cerebral white matter injury of the premature infant – more common than you think. Pediatrics 112, 176–180.CrossRefGoogle Scholar
Volpe, J. J. (2008). Hypoxic-ischemic encephalopathy: neuropathology and pathogenesis. In Neurology of the Newborn (Philadelphia, PA: W.B. Saunders Co.)Google Scholar
Volpe, J. J. (2009). Brain injury in premature infants: a complex amalgam of destructive and developmental disturbances. Lancet Neurol 8, 110–124.CrossRefGoogle ScholarPubMed
Vourc'h, P. and Andres, C. (2004). Oligodendrocyte myelin glycoprotein (OMgp): evolution, structure and function. Brain Res Brain Res Rev 45, 115–124.CrossRefGoogle Scholar
Vouyiouklis, D. A. and Brophy, P. J. (1993). Microtubule-associated protein MAP1B expression precedes the morphological differentiation of oligodendrocytes. J Neurosci Res 35, 257–267.CrossRefGoogle ScholarPubMed
Vouyiouklis, D. A. and Brophy, P. J. (1995). Microtubule-associated proteins in developing oligodendrocytes: transient expression of a MAP2c isoform in oligodendrocyte precursors. J Neurosci Res 42, 803–817.CrossRefGoogle ScholarPubMed
Voyvodic, J. T. (1989). Target size regulates calibre and myelination of sympathetic axons. Nature 342, 430–433.CrossRefGoogle ScholarPubMed
Wakefield, C. L. and Eidelberg, E. (1975). Electron microscopic observations of the delayed effects of spinal cord compression. Exp Neurol 48, 637–646.CrossRefGoogle ScholarPubMed
Wallström, E., Diener, P., Ljungdahl, Ã, Khademi, M., Nilsson, C. G., and Olsson, T. (1996). Memantine abrogates neurological deficits, but not CNS inflammation, in Lewis rat experimental autoimmune encephalomyelitis. J Neurol Sci 137, 89–96.CrossRefGoogle Scholar
Wang Ip, C., Kroner, A., Fischer, S., Berghoff, M., Kobsar, I., Maurer, M., and Martini, R. (2006). Role of immune cells in animal models for inherited peripheral neuropathies. Neuromolecular Med 8, 175–190.Google ScholarPubMed
Wang, H., Tewari, A., Einheber, S., Salzer, J. L., and Melendez-Vasquez, C. V. (2008). Myosin II has distinct functions in PNS and CNS myelin sheath formation. J Cell Biol 182, 1171–1184.CrossRefGoogle ScholarPubMed
Wang, K. C., Kim, J. A., Sivasankaran, R., Segal, R., and He, Z. (2002a). P75 interacts with the Nogo receptor as a co-receptor for Nogo, MAG and OMgp. Nature 420, 74–78.CrossRefGoogle ScholarPubMed
Wang, K. C., Koprivica, V., Kim, J. A., et al. (2002b). Oligodendrocyte-myelin glycoprotein is a Nogo receptor ligand that inhibits neurite outgrowth. Nature 417, 941–944.CrossRefGoogle ScholarPubMed
Wang, S., Sdrulla, A. D., diSibio, G., et al. (1998). Notch receptor activation inhibits oligodendrocyte differentiation. Neuron 21, 63–75.CrossRefGoogle ScholarPubMed
Wang, W., Niekerk, E., Willis, D. E., and Twiss, J. L. (2007). RNA transport and localized protein synthesis in neurological disorders and neural repair. Dev Neurobiol 67, 1166–1182.CrossRefGoogle ScholarPubMed
Wang, W. X., Rajeev, B. W., Stromberg, A. J., et al. (2008). The expression of microRNA miR-107 decreases early in Alzheimer's disease, and may accelerate disease progression through regulation of beta-site amyloid precursor protein-cleaving enzyme 1. J Neurosci 28, 1213–1223.CrossRefGoogle ScholarPubMed
Wanner, I. B., Guerra, N. K., Mahoney, J., et al. (2006a). Role of N-cadherin in Schwann cell precursors of growing nerves. Glia 54, 439–459.CrossRefGoogle ScholarPubMed
Wanner, I. B., Mahoney, J., Jessen, K. R., Wood, P. M., Bates, M., and Bunge, M. B. (2006b). Invariant mantling of growth cones by Schwann cell precursors characterize growing peripheral nerve fronts. Glia 54, 424–438.CrossRefGoogle ScholarPubMed
Warrington, A. E., Asakura, K., Bieber, A. J., et al. (2000). Human monoclonal antibodies reactive to oligodendrocytes promote remyelination in a model of multiple sclerosis. Proc Natl Acad Sci USA 97, 6820–6825.CrossRefGoogle Scholar
Warrington, A. E., Barbarese, E., and Pfeiffer, S. E. (1993). Differential myelinogenic capacity of specific developmental stages of the oligodendrocyte lineage upon transplantation into hypomyelinating hosts. J Neurosci Res 34, 1–13.CrossRefGoogle ScholarPubMed
Warshawsky, I., Rudick, R. A., Staugaitis, S. M., and Natowicz, M. R. (2005). Primary progressive multiple sclerosis as a phenotype of a PLP1 gene mutation. Ann Neurol 58, 470–473.CrossRefGoogle ScholarPubMed
Watanabe, M., Hadzic, T., and Nishiyama, A. (2004). Transient upregulation of Nkx2.2 expression in oligodendrocyte lineage cells during remyelination. Glia 46, 311–322.CrossRefGoogle ScholarPubMed
Watkins, L. R., Hutchinson, M. R., Johnston, I. N., and Maier, S. F. (2005). Glia: novel counter-regulators of opioid analgesia. Trends Neurosci 28, 661–669.CrossRefGoogle ScholarPubMed
Waxman, S. G. (1989). Demyelination in spinal cord injury. J Neurol Sci 91, 1–14.CrossRefGoogle ScholarPubMed
Waxman, S. G. (2006). Axonal conduction and injury in multiple sclerosis: the role of sodium channels. Nat Rev Neurosci 7, 932–941.CrossRefGoogle ScholarPubMed
Waxman, S. G. and Ritchie, J. M. (1993). Molecular dissection of the myelinated axon. Ann Neurol 33, 121–136.CrossRefGoogle ScholarPubMed
Weber, P., Bartsch, U., Rasband, M. N., et al. (1999). Mice deficient for tenascin-R display alterations of the extracellular matrix and decreased axonal conduction velocities in the CNS. J Neurosci 19, 4245–4262.CrossRefGoogle ScholarPubMed
Wegner, M. (2000). Transcriptional control in myelinating glia: the basic recipe. Glia 29, 118–123.3.0.CO;2-Q>CrossRefGoogle ScholarPubMed
Wegner, M., Drolet, D. W., and Rosenfeld, M. G. (1993). Regulation of JC virus by the POU-domain transcription factor Tst-1: implications for progressive multifocal leukoencephalopathy. Proc Natl Acad Sci USA 90, 4743–4747.CrossRefGoogle ScholarPubMed
Wehrle-Haller, B., Koch, M., Baumgartner, S., Spring, J., and Chiquet, M. (1991). Nerve-dependent and -independent tenascin expression in the developing chick limb bud. Development 112, 627–637.Google ScholarPubMed
Weimann, J. M., Charlton, C. A., Brazelton, T. R., Hackman, R. C., and Blau, H. M. (2003). Contribution of transplanted bone marrow cells to Purkinje neurons in human adult brains. Proc Natl Acad Sci USA 100, 2088–2093.CrossRefGoogle ScholarPubMed
Weimbs, T. and Stoffel, W. (1992). Proteolipid protein (PLP) of CNS myelin: positions of free, disulfide-bonded, and fatty acid thioester-linked cysteine residues and implications for the membrane topology of PLP. Biochemistry 31, 12289–12296.CrossRefGoogle ScholarPubMed
Weimbs, T., Low, S. H., Chapin, S. J., and Mostov, K. E. (1997). Apical targeting in polarized epithelial cells: there's more afloat than rafts. Trends Cell Biol 7, 393–399.CrossRefGoogle ScholarPubMed
Weinberg, H. J. and Spencer, P. S. (1975). Studies on the control of myelinogenesis. I. Myelination of regenerating axons after entry into a foreign unmyelinated nerve. J Neurocytol 4, 395–418.CrossRefGoogle ScholarPubMed
Werner, P., Pitt, D., and Raine, C. S. (2000). Glutamate excitotoxicity – a mechanism for axonal damage and oligodendrocyte death in multiple sclerosis? J Neural Transm Suppl (60), 375–385.
Werner, P., Pitt, D., and Raine, C. S. (2001). Multiple sclerosis: altered glutamate homeostasis in lesions correlates with oligodendrocyte and axonal damage. Ann Neurol 50, 169–180.CrossRefGoogle ScholarPubMed
Wiessner, C., Bareyre, F. M., Allegrini, P. R., et al. (2003). Anti-Nogo-A antibody infusion 24 hours after experimental stroke improved behavioral outcome and corticospinal plasticity in normotensive and spontaneously hypertensive rats. J Cereb Blood Flow Metab 23, 154–165.CrossRefGoogle ScholarPubMed
Wight, P. A., Duchala, C. S., Readhead, C., and Macklin, W. B. (1993). A myelin proteolipid protein-LacZ fusion protein is developmentally regulated and targeted to the myelin membrane in transgenic mice. J Cell Biol 123, 443–454.CrossRefGoogle ScholarPubMed
Wigley, R., Hamilton, N., Nishiyama, A., Kirchhoff, F., and Butt, A. M. (2007). Morphological and physiological interactions of NG2-glia with astrocytes and neurons. J Anat 210, 661–670.CrossRefGoogle ScholarPubMed
Wilkins, A. and Compston, A. (2005). Trophic factors attenuate nitric oxide mediated neuronal and axonal injury in vitro: roles and interactions of mitogen-activated protein kinase signalling pathways. J Neurochem. 92, 1487–1496.CrossRefGoogle ScholarPubMed
Wilkins, A. and Scolding, N. (2008). Protecting axons in multiple sclerosis. Mult Scler 14, 1013–1025.CrossRefGoogle ScholarPubMed
Wilkins, A., Chandran, S., and Compston, A. (2001). A role for oligodendrocyte-derived IGF-1 in trophic support of cortical neurons. Glia 36, 48–57.CrossRefGoogle ScholarPubMed
Wilkins, A., Majed, H., Layfield, R., Compston, A., and Chandran, S. (2003). Oligodendrocytes promote neuronal survival and axonal length by distinct intracellular mechanisms: a novel role for oligodendrocyte-derived glial cell line-derived neurotrophic factor. J Neurosci 23, 4967–4974.CrossRefGoogle ScholarPubMed
Wilkinson, D. G., Bhatt, S., Chavrier, P., Bravo, R., and Charnay, P. (1989). Segment-specific expression of a zinc-finger gene in the developing nervous system of the mouse. Nature 337, 461–464.CrossRefGoogle ScholarPubMed
Williams, A., Piaton, G., Aigrot, M. S, et al. (2007a). Semaphorin 3A and 3F: key players in myelin repair in multiple sclerosis? Brain 130, 2554–2465.CrossRefGoogle ScholarPubMed
Williams, A., Piaton, G., and Lubetzki, C. (2007b). Astrocytes – Friends or foes in multiple sclerosis? Glia 55, 1300–1312.CrossRefGoogle ScholarPubMed
Wilson, H., Scolding, N., and Raine, C. (2006). Co-expression of PDGF α receptor and NG2 by oligodendrocyte precursor cells in human CNS and multiple sclerosis lesions. J Neuroimmunol 176, 162–173.CrossRefGoogle ScholarPubMed
Wilson, R. and Brophy, P. J. (1989). Role of the oligodendrocyte cytoskeleton in myelination. J Neurosci Res 22, 439–448.CrossRefGoogle ScholarPubMed
Wilson, S. S., Baetge, E. E., and Stallcup, W. B. (1981). Antisera specific for cell lines with mixed neuronal and glial properties. Dev Biol 83, 146–153.CrossRefGoogle ScholarPubMed
Windrem, M. S., Nunes, M. C., Rashbaum, W. K., et al. (2004). Fetal and adult human oligodendrocyte progenitor cell isolates myelinate the congenitally dysmyelinated brain. Nat Med 10, 93–97.CrossRefGoogle ScholarPubMed
Wingerchuk, D. M. (2004). Neuromyelitis optica: current concepts. Front Biosci 9, 834–840.CrossRefGoogle ScholarPubMed
Wolman, L. (1965). The disturbance of circulation in traumatic paraplegia in acute and late stages: a pathological study. Paraplegia 2, 213–226.Google ScholarPubMed
Wolswijk, G. (1998). Chronic stage multiple sclerosis lesions contain a relatively quiescent population of oligodendrocyte precursor cells. J Neurosci 18, 601–609.CrossRefGoogle ScholarPubMed
Wolswijk, G. (2002). Oligodendrocyte precursor cells in the demyelinated multiple sclerosis spinal cord. Brain 125, 338–349.CrossRefGoogle ScholarPubMed
Wolswijk, G. and Noble, M. (1989). Identification of an adult-specific glial progenitor cell. Development 105, 387–400.Google ScholarPubMed
Wolswijk, G., Munro, P. M. G., Riddle, P. N., and Noble, M. (1991). Origin, growth factor responses, and ultrastructural characteristics of an adult-specific glial progenitor cell. Ann New York Acad Sci 633, 502–504.CrossRefGoogle ScholarPubMed
Wong, S. T., Henley, J. R., Kanning, K. C., Huang, K. H., Bothwell, M., and Poo, M. M. (2002). A p75(NTR) and Nogo receptor complex mediates repulsive signaling by myelin-associated glycoprotein. Nat Neurosci 5, 1302–1308.CrossRefGoogle ScholarPubMed
Wood, P. M., Schachner, M., and Bunge, R. P. (1990). Inhibition of Schwann cell myelination in vitro by antibody to the L1 adhesion molecule. J Neurosci 10, 3635–3645.CrossRefGoogle ScholarPubMed
Woodhoo, A. and Sommer, L. (2008). Development of the Schwann cell lineage: from the neural crest to the myelinated nerve. Glia 56, 1481–1490.CrossRefGoogle ScholarPubMed
Woodhoo, A., Sahni, V., Gilson, J., et al. (2007). Schwann cell precursors: a favourable cell for myelin repair in the central nervous system. Brain 130, 2175–2185.CrossRefGoogle ScholarPubMed
Woodruff, R. H., Fruttiger, M., Richardson, W. D., and Franklin, R. J. (2004). Platelet-derived growth factor regulates oligodendrocyte progenitor numbers in adult CNS and their response following CNS demyelination. Mol Cell Neurosci 25, 252–262.CrossRefGoogle ScholarPubMed
Woodward, L. J., Edgin, J. O., Thompson, D., and Inder, T. E. (2005). Object working memory deficits predicted by early brain injury and development in the preterm infant. Brain 128, 2578–2587.CrossRefGoogle ScholarPubMed
Wosik, K., Antel, J., Kuhlmann, T., Brück, W., Massie, B., and Nalbantoglu, J. (2003). Oligodendrocyte injury in multiple sclerosis: a role for p53. J Neurochem 85, 635–644.CrossRefGoogle ScholarPubMed
Wrathall, J. R., Li, W., and Hudson, L. D. (1998). Myelin gene expression after experimental contusive spinal cord injury. J Neurosci 18, 8780–8793.CrossRefGoogle ScholarPubMed
Wren, D., Wolswijk, G., and Noble, M. (1992). In vitro analysis of the origin and maintenance of O-2A(adult) progenitor cells. J Cell Biol 116, 167–176.CrossRefGoogle ScholarPubMed
Wu, Q., Miller, R. H., Ransohoff, R. M., Robinson, S., Bu, J., and Nishiyama, A. (2000). Elevated levels of the chemokine GRO-1 correlate with elevated oligodendrocyte progenitor proliferation in the jimpy mutant. J Neurosci 20, 2609–2617.CrossRefGoogle ScholarPubMed
Wucherpfennig, K. W., Newcombe, J., Li, H., Keddy, C., Cuzner, M. L., and Hafler, D. A. (1992). δ T-cell receptor repertoire in acute multiple sclerosis lesions. Proc Natl Acad Sci USA 89, 4588–4592.CrossRefGoogle ScholarPubMed
Xu, W., Fazekas, G., Hara, H., and Tabira, T. (2005). Mechanism of natural killer (NK) cell regulatory role in experimental autoimmune encephalomyelitis. J Neuroimmunol 163, 24–30.CrossRefGoogle ScholarPubMed
Xu, W., Shy, M., Kamholz, J., et al. (2001). Mutations in the cytoplasmic domain of P0 reveal a role for PKC-mediated phosphorylation in adhesion and myelination. J Cell Biol 155, 439–446.CrossRefGoogle ScholarPubMed
Yamada, K., Watanabe, M., Shibata, T., Nagashima, M., Tanaka, K., and Inoue, Y. (1998). Glutamate transporter GLT-1 is transiently localized on growing axons of the mouse spinal cord before establishing astrocytic expression. J Neurosci 18, 5706–5713.CrossRefGoogle ScholarPubMed
Yamada, M., Mizuguchi, M., Otsuka, N., Ikeda, K., and Takahashi, H. (1997). Ultrastructural localization of CD38 immunoreactivity in rat brain. Brain Res 756, 52–60.CrossRefGoogle ScholarPubMed
Yamamoto, S., Yamamoto, N., Kitamura, T., Nakamura, K., and Nakafuku, M. (2001). Proliferation of parenchymal neural progenitors in response to injury in the adult rat spinal cord. Exp Neurol 172, 115–127.CrossRefGoogle ScholarPubMed
Yamamoto, Y., Mizuno, R., Nishimura, T., et al. (1994). Cloning and expression of myelin-associated oligodendrocytic basic protein. A novel basic protein constituting the central nervous system myelin. J Biol Chem 269, 31725–31730.Google ScholarPubMed
Yan, P., Liu, N., Kim, G. M., et al. (2003). Expression of the type 1 and type 2 receptors for tumor necrosis factor after traumatic spinal cord injury in adult rats. Exp Neurol 183, 286–297.CrossRefGoogle ScholarPubMed
Yang, D. P., Zhang, D. P., Mak, K. S., Bonder, D. E., Pomeroy, S. L., and Kim, H. A. (2008). Schwann cell proliferation during Wallerian degeneration is not necessary for regeneration and remyelination of the peripheral nerves: axon-dependent removal of newly generated Schwann cells by apoptosis. Mol Cell Neurosci 38, 80–88.CrossRefGoogle Scholar
Yang, Y., Ogawa, Y., Hedstrom, K. L., and Rasband, M. N. (2007). betaIV spectrin is recruited to axon initial segments and nodes of Ranvier by ankyrinG. J Cell Biol 176, 509–519.CrossRefGoogle ScholarPubMed
Ye, M., Chen, S., Wang, X., et al. (2005). Glial cell line-derived neurotrophic factor in bone marrow stromal cells of rat. NeuroReport 16, 581–584.CrossRefGoogle ScholarPubMed
Ye, P., Carson, J., and D'Ercole, A. J. (1995). In vivo actions of insulin-like growth factor-I (IGF-I) on brain myelination: studies of IGF-I and IGF binding protein-1 (IGFBP-1) transgenic mice. J Neurosci 15, 7344–7356.CrossRefGoogle ScholarPubMed
Yednock, T. A., Cannon, C., Fritz, L. C., Sanchez-Madrid, F., Steinman, L., and Karin, N. (1992). Prevention of experimental autoimmune encephalomyelitis by antibodies against alpha 4 beta 1 integrin. Nature 356, 63–66.CrossRefGoogle ScholarPubMed
Yin, X., Baek, R. C., Kirschner, D. A., et al. (2006). Evolution of a neuroprotective function of central nervous system myelin. J Cell Biol 172, 469–478.CrossRefGoogle ScholarPubMed
Yin, X., Crawford, T. O., Griffin, J. W., et al. (1998). Myelin-associated glycoprotein is a myelin signal that modulates the caliber of myelinated axons. J Neurosci 18, 1953–1962.CrossRefGoogle ScholarPubMed
Yin, X., Peterson, J., Gravel, M., Braun, P. E., and Trapp, B. D. (1997). CNP overexpression induces aberrant oligodendrocyte membranes and inhibits MBP accumulation and myelin compaction. J Neurosci Res 50, 238–247.3.0.CO;2-4>CrossRefGoogle ScholarPubMed
Yiu, G. and He, Z. (2006). Glial inhibition of CNS axon regeneration. Nat Rev Neurosci 7, 617–627.CrossRefGoogle ScholarPubMed
Yool, D. A., Edgar, J. M., Montague, P., and Malcolm, S. (2000). The proteolipid protein gene and myelin disorders in man and animal models. Hum Mol Genet 9, 987–992.CrossRefGoogle ScholarPubMed
Yoon, B. H., Kim, C. J., Romero, R., et al. (1997). Experimentally induced intrauterine infection causes fetal brain white matter lesions in rabbits. Am J Obstet Gynecol 177, 797–802.CrossRefGoogle ScholarPubMed
Yoshida, M. and Colman, D. R. (1996). Parallel evolution and coexpression of the proteolipid proteins and protein zero in vertebrate myelin. Neuron 16, 1115–1126.CrossRefGoogle ScholarPubMed
Yoshino, J. E., Mason, P. W., and DeVries, G. H. (1987). Developmental changes in myelin-induced proliferation of cultured Schwann cells. J Cell Biol 104, 655–660.CrossRefGoogle ScholarPubMed
Yoshioka, A., Hardy, M., Younkin, D. P., Grinspan, J. B., Stern, J. L., and Pleasure, D. (1995). α-Amino-3-hydroxy-5-methyl-4-isoxazolepropionate (AMPA) receptors mediate excitotoxicity in the oligodendroglial lineage. J Neurochem 64, 2442–2448.CrossRefGoogle ScholarPubMed
Yoshioka, A., Ikegaki, N., Williams, M., and Pleasure, D. (1996). Expression of N-methyl-d-aspartate (NMDA) and non-NMDA glutamate receptor genes in neuroblastoma, medulloblastoma, and other cell lines. J Neurosci Res 46, 164–172.3.0.CO;2-F>CrossRefGoogle Scholar
Young, E. A., Fowler, C. D., Kidd, G. J., et al. (2008). Imaging correlates of decreased axonal Na+/K+ ATPase in chronic MS lesions. Ann Neurol 63, 428–435.CrossRefGoogle Scholar
Yousry, T. A., Major, E. O., Ryschkewitsch, C., et al. (2006). Evaluation of patients treated with natalizumab for progressive multifocal leukoencephalopathy. N Engl J Med 354, 924–933.CrossRefGoogle ScholarPubMed
Yu, J., Vodyanik, M. A., Smuga-Otto, K., et al. (2007). Induced pluripotent stem cell lines derived from human somatic cells. Science 318, 1917–1920.CrossRefGoogle ScholarPubMed
Yu, W. M., Feltri, M. L., Wrabetz, L., Strickland, S., and Chen, Z. L. (2005). Schwann cell-specific ablation of laminin gamma1 causes apoptosis and prevents proliferation. J Neurosci 25, 4463–4472.CrossRefGoogle ScholarPubMed
Yuan, X., Eisen, A. M., McBain, C. J., and Gallo, V. (1998). A role for glutamate and its receptors in the regulation of oligodendrocyte development in cerebellar tissue slices. Development 125, 2901–2914.Google ScholarPubMed
Zai, L. J. and Wrathall, J. R. (2005). Cell proliferation and replacement following contusive spinal cord injury. Glia 50, 247–257.CrossRefGoogle ScholarPubMed
Zajicek, J., Wing, M., Skepper, J., and Compston, A. (1995). Human oligodendrocytes are not sensitive to complement. A study of CD59 expression in the human central nervous system. Lab Invest 73, 128–138.Google Scholar
Zajicek, J. P., Wing, M., Scolding, N. J., and Compston, D. A. (1992). Interactions between oligodendrocytes and microglia. A major role for complement and tumour necrosis factor in oligodendrocyte adherence and killing. Brain 115, 1611–1631.CrossRefGoogle Scholar
Zalc, B. (2006). The acquisition of myelin: a success story. Novartis Found Symp 276, 15–21.Google ScholarPubMed
Zalc, B., Goujet, D., and Colman, D. (2008). The origin of the myelination program in vertebrates. Curr Biol 18, R511–R512.CrossRefGoogle ScholarPubMed
Zang, Y. C., Li, S., Rivera, V. M., et al. (2004). Increased CD8(+) cytotoxic T cell responses to myelin basic protein in multiple sclerosis. J Immunol 172, 5120–5127.CrossRefGoogle ScholarPubMed
Zappia, E., Casazza, S., Pedemonte, E., et al. (2005). Mesenchymal stem cells ameliorate experimental autoimmune encephalomyelitis inducing T-cell anergy. Blood 106, 1755–1761.CrossRefGoogle ScholarPubMed
Zeck-Kapp, G., Kroegel, C., Riede, U. N., and Kapp, A. (1995). Mechanisms of human eosinophil activation by complement protein C5a and platelet-activating factor: similar functional responses are accompanied by different morphologic alterations. Allergy 50, 34–47.CrossRefGoogle ScholarPubMed
Zeger, M., Popken, G., Zhang, J., et al. (2007). Insulin-like growth factor type 1 receptor signaling in the cells of oligodendrocyte lineage is required for normal in vivo oligodendrocyte development and myelination. Glia 55, 400–411.CrossRefGoogle ScholarPubMed
Zeis, T., Graumann, U., Reynolds, R., and Schaeren-Wiemers, N. (2008a). Normal-appearing white matter in multiple sclerosis is in a subtle balance between inflammation and neuroprotection. Brain 131, 288–303.CrossRefGoogle Scholar
Zeis, T., Kinter, J., Herrero-Herranz, E., Weissert, R., and Schaeren-Wiemers, N. (2008b). Gene expression analysis of normal appearing brain tissue in an animal model for multiple sclerosis revealed grey matter alterations, but only minor white matter changes. J Neuroimmunol 205, 10–19.CrossRefGoogle Scholar
Zeller, N. K., Hunkeler, M. J., Campagnoni, A. T., Sprague, J., and Lazzarini, R. A. (1984). Characterization of mouse myelin basic protein messenger RNAs with a myelin basic protein cDNA clone. Proc Natl Acad Sci USA 81, 18–22.CrossRefGoogle ScholarPubMed
Zerangue, N. and Kavanaugh, M. P. (1996). Flux coupling in a neuronal glutamate transporter. Nature 383, 634–637.CrossRefGoogle Scholar
Zhang, J., Li, Y., Chen, J., et al. (2005). Human bone marrow stromal cell treatment improves neurological functional recovery in EAE mice. Exp Neurol 195, 16–26.CrossRefGoogle ScholarPubMed
Zhang, J., Li, Y., Lu, M., et al. (2006). Bone marrow stromal cells reduce axonal loss in experimental autoimmune encephalomyelitis mice. J Neurosci Res 84, 587–595.CrossRefGoogle ScholarPubMed
Zhang, S. C., Ge, B., and Duncan, I. D. (1999). Adult brain retains the potential to generate oligodendroglial progenitors with extensive myelination capacity. Proc Natl Acad Sci USA 96, 4089–4094.CrossRefGoogle ScholarPubMed
Zhang, Y., Wang, H., Li, J., et al. (2006). Intracellular zinc release and ERK phosphorylation are required upstream of 12-lipoxygenase activation in peroxynitrite toxicity to mature rat oligodendrocytes. J Biol Chem 281, 9460–9470.CrossRefGoogle ScholarPubMed
Zheng, B., Atwal, J., Ho, C., et al. (2005). Genetic deletion of the Nogo receptor does not reduce neurite inhibition in vitro or promote corticospinal tract regeneration in vivo. Proc Natl Acad Sci USA 102, 1205–1210.CrossRefGoogle ScholarPubMed
Zheng, B., Ho, C., Li, S., Keirstead, H., Steward, O., and Tessier-Lavigne, M. (2003). Lack of enhanced spinal regeneration in Nogo-deficient mice. Neuron 38, 213–224.CrossRefGoogle ScholarPubMed
Zhou, D., Srivastava, R., Nessler, S., et al. (2006). Identification of a pathogenic antibody response to native myelin oligodendrocyte glycoprotein in multiple sclerosis. Proc Natl Acad Sci USA 103, 19057–19062.CrossRefGoogle ScholarPubMed
Zhou, Q. and Anderson, D. J. (2002). The bHLH transcription factors OLIG2 and OLIG1 couple neuronal and glial subtype specification. Cell 109, 61–73.CrossRefGoogle ScholarPubMed
Zhou, Q., Choi, G., and Anderson, D. J. (2001). The bHLH transcription factor Olig2 promotes oligodendrocyte differentiation in collaboration with Nkx2.2. Neuron 31, 791–807.CrossRefGoogle ScholarPubMed
Zhou, Q., Wang, S., and Anderson, D. J. (2000). Identification of a novel family of oligodendrocyte lineage-specific basic helix-loop-helix transcription factors. Neuron 25, 331–343.CrossRefGoogle ScholarPubMed
Zhu, X., Bergles, D. E., and Nishiyama, A. (2008). NG2 cells generate both oligodendrocytes and gray matter astrocytes. Development 135, 145–157.CrossRefGoogle ScholarPubMed
Ziak, D., Chvatal, A., and Sykova, E. (1998). Glutamate-, kainate- and NMDA-evoked membrane currents in identified glial cells in rat spinal cord slice. Physiol Res 47, 365–375.Google ScholarPubMed
Ziskin, J. L., Nishiyama, A., Rubio, M., Fukaya, M., and Bergles, D. E. (2007). Vesicular release of glutamate from unmyelinated axons in white matter. Nat Neurosci 10, 321–330.CrossRefGoogle ScholarPubMed
Zuo, J., Ferguson, T. A., Hernandez, Y. J., Stetler-Stevenson, W. G., and Muir, D. (1998). Neuronal matrix metalloproteinase-2 degrades and inactivates a neurite-inhibiting chondroitin sulfate proteoglycan. J Neurosci 18, 5203–5211.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Edited by Patricia Armati, University of Sydney, Emily Mathey, University of Sydney
  • Book: The Biology of Oligodendrocytes
  • Online publication: 05 August 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511782121.012
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Edited by Patricia Armati, University of Sydney, Emily Mathey, University of Sydney
  • Book: The Biology of Oligodendrocytes
  • Online publication: 05 August 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511782121.012
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Edited by Patricia Armati, University of Sydney, Emily Mathey, University of Sydney
  • Book: The Biology of Oligodendrocytes
  • Online publication: 05 August 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511782121.012
Available formats
×