Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Motor learning drives dynamic patterns of intermittent myelination on learning-activated axons

Abstract

Myelin plasticity occurs when newly formed and pre-existing oligodendrocytes remodel existing patterns of myelination. Myelin remodeling occurs in response to changes in neuronal activity and is required for learning and memory. However, the link between behavior-induced neuronal activity and circuit-specific changes in myelination remains unclear. Using longitudinal in vivo two-photon imaging and targeted labeling of learning-activated neurons in mice, we explore how the pattern of intermittent myelination is altered on individual cortical axons during learning of a dexterous reach task. We show that behavior-induced myelin plasticity is targeted to learning-activated axons and occurs in a staged response across cortical layers in the mouse primary motor cortex. During learning, myelin sheaths retract, which results in lengthening of nodes of Ranvier. Following motor learning, addition of newly formed myelin sheaths increases the number of continuous stretches of myelination. Computational modeling suggests that motor learning-induced myelin plasticity initially slows and subsequently increases axonal conduction speed. Finally, we show that both the magnitude and timing of nodal and myelin dynamics correlate with improvement of behavioral performance during motor learning. Thus, learning-induced and circuit-specific myelination changes may contribute to information encoding in neural circuits during motor learning.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Learning-induced sheath retraction modifies node length.
Fig. 2: Myelin and nodal plasticity occur in distinct phases in response to learning.
Fig. 3: Learning-induced changes to myelin occur at specific locations along axons.
Fig. 4: Learning induces myelin and nodal dynamics broadly across cortical layers.
Fig. 5: Modeled effect of learning-induced sheath and nodal dynamics on conduction.
Fig. 6: Learning-activated axons exhibit heightened myelin dynamics.
Fig. 7: The extent and timing of myelin dynamics is correlated with behavioral performance.

Similar content being viewed by others

Data availability

All data that support the findings, tools and reagents will be shared on an unrestricted basis; requests should be directed to the corresponding author.

Code availability

Code for analysis associated with the manuscript is available at https://github.com/EthanHughesLab/Bacmeisteretal2022.

References

  1. Lillie, R. S. Factors affecting transmission and recovery in the passive iron nerve model. J. Gen. Physiol. 7, 473–507 (1925).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Tomassy, G. S. et al. Distinct profiles of myelin distribution along single axons of pyramidal neurons in the neocortex. Science 344, 319–324 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Yakovlev, P. & Lecours, A. The myelogenetic cycles of regional maturation of the brain. In Regional Development of the Brain in Early Life (ed. Minkowski, A.) 3–70 (Blackwell Scientific Publications, 1967).

  4. Gibson, E. M. et al. Neuronal activity promotes oligodendrogenesis and adaptive myelination in the mammalian brain. Science 344, 1252304 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  5. Hughes, E. G., Orthmann-Murphy, J. L., Langseth, A. J. & Bergles, D. E. Myelin remodeling through experience-dependent oligodendrogenesis in the adult somatosensory cortex. Nat. Neurosci. 21, 696–706 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. McKenzie, I. A. et al. Motor skill learning requires active central myelination. Science 346, 318–322 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Pan, S., Mayoral, S. R., Choi, H. S., Chan, J. R. & Kheirbek, M. A. Preservation of a remote fear memory requires new myelin formation. Nat. Neurosci. 23, 487–499 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Steadman, P. E. et al. Disruption of oligodendrogenesis impairs memory consolidation in adult mice. Neuron 105, 150–164.e6 (2020).

    Article  CAS  PubMed  Google Scholar 

  9. Hines, J. H., Ravanelli, A. M., Schwindt, R., Scott, E. K. & Appel, B. Neuronal activity biases axon selection for myelination in vivo. Nat. Neurosci. 18, 683–689 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Mensch, S. et al. Synaptic vesicle release regulates myelin sheath number of individual oligodendrocytes in vivo. Nat. Neurosci. 18, 628–630 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Mitew, S. et al. Pharmacogenetic stimulation of neuronal activity increases myelination in an axon-specific manner. Nat. Commun. 9, 306 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  12. Baraban, M., Koudelka, S. & Lyons, D. A. Ca2+ activity signatures of myelin sheath formation and growth in vivo. Nat. Neurosci. 21, 19–23 (2018).

    Article  CAS  PubMed  Google Scholar 

  13. Krasnow, A. M., Ford, M. C., Valdivia, L. E., Wilson, S. W. & Attwell, D. Regulation of developing myelin sheath elongation by oligodendrocyte calcium transients in vivo. Nat. Neurosci. 21, 24–28 (2018).

    Article  CAS  PubMed  Google Scholar 

  14. Seidl, A. H., Rubel, E. W. & Harris, D. M. Mechanisms for adjusting interaural time differences to achieve binaural coincidence detection. J. Neurosci. 30, 70–80 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Ford, M. C. et al. Tuning of Ranvier node and internode properties in myelinated axons to adjust action potential timing. Nat. Commun. 6, 8073 (2015).

    Article  CAS  PubMed  Google Scholar 

  16. Arancibia-Cárcamo, I. L. et al. Node of Ranvier length as a potential regulator of myelinated axon conduction speed. eLife 6, e23329 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  17. Salami, M., Itami, C., Tsumoto, T. & Kimura, F. Change of conduction velocity by regional myelination yields constant latency irrespective of distance between thalamus and cortex. Proc. Natl Acad. Sci USA 100, 6174–6179 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Auer, F., Vagionitis, S. & Czopka, T. Evidence for myelin sheath remodeling in the CNS revealed by in vivo imaging. Curr. Biol. 28, 549–559.e3 (2018).

    Article  CAS  PubMed  Google Scholar 

  19. Bacmeister, C. M. et al. Motor learning promotes remyelination via new and surviving oligodendrocytes. Nat. Neurosci. 23, 819–831 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Cullen, C. L. et al. Periaxonal and nodal plasticities modulate action potential conduction in the adult mouse brain. Cell Rep. 34, 108641 (2021).

    Article  CAS  PubMed  Google Scholar 

  21. Yang, S. M., Michel, K., Jokhi, V., Nedivi, E. & Arlotta, P. Neuron class–specific responses govern adaptive myelin remodeling in the neocortex. Science 370, eabd2109 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Bishop, D. et al. Near-infrared branding efficiently correlates light and electron microscopy. Nat. Methods 8, 568–570 (2011).

    Article  CAS  PubMed  Google Scholar 

  23. Bostock, H. & Sears, T. A. The internodal axon membrane: electrical excitability and continuous conduction in segmental demyelination. J. Physiol. 280, 273–301 (1978).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Foster, R. E., Whalen, C. C. & Waxman, S. G. Reorganization of the axon membrane in demyelinated peripheral nerve fibers: morphological evidence. Science 210, 661–663 (1980).

    Article  CAS  PubMed  Google Scholar 

  25. Rasminsky, M. & Sears, T. A. Internodal conduction in undissected demyelinated nerve fibres. J. Physiol. 227, 323–350 (1972).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Coggan, J. S., Prescott, S. A., Bartol, T. M. & Sejnowski, T. J. Imbalance of ionic conductances contributes to diverse symptoms of demyelination. Proc. Natl Acad. Sci. USA 107, 20602–20609 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Ye, L. et al. Wiring and molecular features of prefrontal ensembles representing distinct experiences. Cell 165, 1776–1788 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Papale, A. E. & Hooks, B. M. Circuit changes in motor cortex during motor skill learning. Neuroscience 368, 283–297 (2018).

    Article  CAS  PubMed  Google Scholar 

  29. Young, K. M. et al. Oligodendrocyte dynamics in the healthy adult CNS: evidence for myelin remodeling. Neuron 77, 873–885 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Hill, R. A., Li, A. M. & Grutzendler, J. Lifelong cortical myelin plasticity and age-related degeneration in the live mammalian brain. Nat. Neurosci. 21, 683–695 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Koudelka, S. et al. Individual neuronal subtypes exhibit diversity in CNS myelination mediated by synaptic vesicle release. Curr. Biol. 26, 1447–1455 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Bechler, M. E., Byrne, L. & ffrench-Constant, C. CNS myelin sheath lengths are an intrinsic property of oligodendrocytes. Curr. Biol. 25, 2411–2416 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Swire, M. et al. Oligodendrocyte HCN2 channels regulate myelin sheath length. J. Neurosci. 41, 7954–7964 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Chong, S. Y. C. et al. Neurite outgrowth inhibitor Nogo-A establishes spatial segregation and extent of oligodendrocyte myelination. Proc. Natl Acad. Sci. USA 109, 1299–1304 (2012).

    Article  CAS  PubMed  Google Scholar 

  35. Dhawale, A. K., Smith, M. A. & Ölveczky, B. P. The role of variability in motor learning. Annu. Rev. Neurosci. 40, 479–498 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Feldman, D. E. The spike-timing dependence of plasticity. Neuron 75, 556–571 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Fields, R. D. A new mechanism of nervous system plasticity: activity-dependent myelination. Nat. Rev. Neurosci. 16, 756–767 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Egger, R. et al. Local axonal conduction shapes the spatiotemporal properties of neural sequences. Cell 183, 537–548.e12 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. de Waegh, S. M., Lee, V. M.-Y. & Brady, S. T. Local modulation of neurofilament phosphorylation, axonal caliber, and slow axonal transport by myelinating Schwann cells. Cell 68, 451–463 (1992).

    Article  PubMed  Google Scholar 

  40. Fünfschilling, U. et al. Glycolytic oligodendrocytes maintain myelin and long-term axonal integrity. Nature 485, 517–521 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Moore, S. et al. A role of oligodendrocytes in information processing. Nat. Commun. 11, 5497 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Harris, J. J. & Attwell, D. The energetics of CNS white matter. J. Neurosci. 32, 356–371 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Lee, Y. et al. Oligodendroglia metabolically support axons and contribute to neurodegeneration. Nature 487, 443–448 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Brown, A. M., Tekkök, S. B. & Ransom, B. R. Glycogen regulation and functional role in mouse white matter. J. Physiol. 549, 501–512 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Trevisiol, A. et al. Structural myelin defects are associated with low axonal ATP levels but rapid recovery from energy deprivation in a mouse model of spastic paraplegia. PLoS Biol. 18, e3000943 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Stedehouder, J. et al. Fast-spiking parvalbumin interneurons are frequently myelinated in the cerebral cortex of mice and humans. Cereb. Cortex 27, 5001–5013 (2017).

    Article  CAS  PubMed  Google Scholar 

  47. Sasaki, T., Matsuki, N. & Ikegaya, Y. Action-potential modulation during axonal conduction. Science 331, 599–601 (2011).

    Article  CAS  PubMed  Google Scholar 

  48. Kalish, B. T. et al. Single-nucleus RNA sequencing of mouse auditory cortex reveals critical period triggers and brakes. Proc. Natl Acad. Sci. USA 117, 11744–11752 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Cichon, J. & Gan, W.-B. Branch-specific dendritic Ca2+ spikes cause persistent synaptic plasticity. Nature 520, 180–185 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Fu, M., Yu, X., Lu, J. & Zuo, Y. Repetitive motor learning induces coordinated formation of clustered dendritic spines in vivo. Nature 483, 92–95 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Stedehouder, J., Brizee, D., Shpak, G. & Kushner, S. A. Activity-dependent myelination of parvalbumin interneurons mediated by axonal morphological plasticity. J. Neurosci. 38, 3631–3642 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Bramow, S. et al. Demyelination versus remyelination in progressive multiple sclerosis. Brain 133, 2983–2998 (2010).

    Article  PubMed  Google Scholar 

  53. Patrikios, P. et al. Remyelination is extensive in a subset of multiple sclerosis patients. Brain 129, 3165–3172 (2006).

    Article  PubMed  Google Scholar 

  54. Goldschmidt, T., Antel, J., König, F. B., Brück, W. & Kuhlmann, T. Remyelination capacity of the MS brain decreases with disease chronicity. Neurology 72, 1914–1921 (2009).

    Article  CAS  PubMed  Google Scholar 

  55. Snaidero, N. et al. Myelin replacement triggered by single-cell demyelination in mouse cortex. Nat. Commun. 11, 4901 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Orthmann-Murphy, J. et al. Remyelination alters the pattern of myelin in the cerebral cortex. eLife 9, e56621 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Hughes, E. G., Kang, S. H., Fukaya, M. & Bergles, D. E. Oligodendrocyte progenitors balance growth with self-repulsion to achieve homeostasis in the adult brain. Nat. Neurosci. 16, 668–676 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Xu, T. et al. Rapid formation and selective stabilization of synapses for enduring motor memories. Nature 462, 915–919 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Hines, M. L. & Carnevale, N. T. The NEURON Simulation Environment. Neural Comput. 9, 1179–1209 (1997).

    Article  CAS  PubMed  Google Scholar 

  60. Cohen, C. C. H. et al. Saltatory conduction along myelinated axons involves a periaxonal nanocircuit. Cell 180, 311–322.e15 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Hu, W. et al. Distinct contributions of Nav1.6 and Nav1.2 in action potential initiation and backpropagation. Nat. Neurosci. 12, 996–1002 (2009).

    Article  CAS  PubMed  Google Scholar 

  62. Mainen, Z. F., Joerges, J., Huguenard, J. R. & Sejnowski, T. J. A model of spike initiation in neocortical pyramidal neurons. Neuron 15, 1427–1439 (1995).

    Article  CAS  PubMed  Google Scholar 

  63. Hodgkin, A. L. & Huxley, A. F. A quantitative description of membrane current and its application to conduction and excitation in nerve. J. Physiol. 117, 500–544 (1952).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Huguenard, J. R., Hamill, O. P. & Prince, D. A. Developmental changes in Na+ conductances in rat neocortical neurons: appearance of a slowly inactivating component. J. Neurophysiol. 59, 778–795 (1988).

    Article  CAS  PubMed  Google Scholar 

  65. Hamill, O. P., Huguenard, J. R. & Prince, D. A. Patch-clamp studies of voltage-gated currents in identified neurons of the rat cerebral cortex. Cereb. Cortex 1, 48–61 (1991).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank M. Hall for machining expertise, H. Barr for the conversation that reinitiated interest in labeling activated axons, and past and current members of the Hughes and Welle labs and the CU Anschutz Myelin Group for discussions. M.A.T. is supported by the National Institutes of Health Institutional Neuroscience Graduate Training Grant (5T32NS099042-17). Funding was provided by the Boettcher Foundation Webb-Waring Biomedical Research Award, the Whitehall Foundation and the National Multiple Sclerosis Society (RG-1701–26733) and NINDS (NS115975, NS125230) to E.G.H.

Author information

Authors and Affiliations

Authors

Contributions

E.G.H. and C.M.B. conceived the project. C.M.B. designed and performed experiments, analyzed data and generated all figures. R.H. performed all viral injections and performed surgeries. L.A.O. performed experiments and analyzed data for Fig. 7 and Extended Data Figs. 2 and 3. M.A.T. performed surgeries. L.C. performed experiments and additional imaging analyses for Figs. 4, 6 and 7 and Extended Data Figs. 2 and 3. A.R.C. provided technical support for all mouse lines. A.P.-P. supervised the computational modulating and E.G.H. supervised the project. C.M.B. and E.G.H. wrote the manuscript with input from other authors.

Corresponding author

Correspondence to Ethan G. Hughes.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Neuroscience thanks Michelle Monje and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Near-infrared branding identifies the same oligodendrocytes and myelin sheaths in longitudinally, in vivo imaged areas and posthoc stained tissue.

a, The same field of view imaged in vivo (‘Live Tissue’, left) and in fixed, sectioned, and stained tissue (‘Fixed Tissue’, right). Patterns of cell bodies (examples outlined in white dotted lines) were maintained across live and processed tissue. Note the new oligodendrocyte generated at 60d and delineated with a white arrowhead. b, A newly generated oligodendrocyte in vivo (‘Live Tissue’, left) and fixed, sectioned tissue (‘Fixed Tissue’, right) stained for oligodendrocytes and myelin (GFP, green) and axons (NFH, magenta). Note the same T-junction across live and fixed samples is marked with the white arrowhead.

Extended Data Fig. 2 Learning trajectory of mice engaging in forelimb reach training.

a, Illustration of forelimb reach training box environment. Mice learn to reach by extending their left hand through a slit in a plexiglass box to grab a pellet and return it to their mouth. b, Imaging and training timelines for untrained (top) and learning (bottom) mice. c, The majority of mice learn to perform the forelimb reach task (learners, green). Learners improve their success rate gradually over the course of seven days of training. In contrast, non-learners maintain a low success rate across time and do not attain higher than a 15% success rate at the end of the training regimen (purple). d, Successful learners of the task perform significantly better on the last day of training (Paired Student’s t-test, t(15)=11.72, p < 0.0001) and achieve higher than 15% success rate, in contrast to non-learners which do not improve significantly and do not attain above 15% success on the last day of training. Mice were excluded from data analysis if they did not succeed in at least 10% of reaches across the seven days of training (n = 2 mice). e, f, Breakdown of all mice included in each of the figures, where each dot is a unique color that represents a single mouse and is consistent across graphs. Neither success rate (e) nor change in success rate (f) differ across figures (REML with post hoc Tukey’s HSD). Bars and error bars represent mean ± s.e.m unless otherwise noted. For detailed statistics, see Supplementary Table 3.

Extended Data Fig. 3 Identifying and tracing nodes and axons across live and fixed tissue.

a, The same node in live imaged (‘Live Tissue’, left) and fixed, sectioned tissue (‘Fixed Tissue’, right) stained for oligodendrocytes and myelin (GFP, green), sodium channels (NaV, magenta), and axons (NFH, yellow). White arrow points to node across time and in fixed tissue. b,c, 90.90 ± 3.73% of nodes visualized in vivo possess NaV staining characteristic of nodes of Ranvier (b, c), comparable to values of nodes identified fixed tissue of age-matched controls (c; random sample, nmice = 6; longitudinally imaged, nmice = 3). d, No difference in the proportion of gaps in myelination (larger than 3 microns) with and without appreciable sodium channel distributions in fixed tissue of age-matched controls (nmice = 8; Paired Student’s t-test). e, Representative in vivo imaging and post hoc immunostaining of a lengthening node with sodium channels (left) and without appreciable sodium channels (right). f, The same node confirmed with NaV staining in (a) is identified in live imaging by a marked decrease in autofluorescence between two GFP-labeled myelin paranodes. g, Axon morphology is accurately identified in vivo as confirmed by post hoc immunostaining of the same tissue. A lengthening node in vivo (‘Live Tissue’, left) and fixed, sectioned tissue (‘Fixed Tissue’, middle). Axon trajectory is identified using the shape of the myelin sheath prior to node lengthening (‘Live axon morphology’, magenta) and confirmed using post hoc immunostaining of the same sheaths for axons (NFH, magenta) and myelin (GFP, green) and reconstructions of the axons (‘Fixed axon morphology’, magenta). In vivo node morphology (‘In vivo nodal morphology’, top right) is reconstructed and confirmed using post hoc immunostaining and morphological reconstruction (‘Fixed nodal morphology’, bottom right). h, Sheath length and change in sheath length is determined by aligning fields of view using fiduciary marks (for example cell soma) which extend across the entire duration of the study, including pre-existing oligodendrocytes, which maintain their position throughout the course of imaging (left). Change in sheath length is mirrored by a change in paranode position (right). To determine the change in sheath length, the distance between the initial and final paranodal position is traced using Simple Neurite Tracer and overlayed stacks. Bars and error bars represent mean ± s.e.m unless otherwise noted. For detailed statistics, see Supplementary Table 3.

Extended Data Fig. 4 Sheath retraction, but not sheath growth, is affected by learning a new skill.

a, Sheath length is similar across sheaths with 2, 1, or 0 neighbors. b, Sheaths of many ages initiate sheath dynamics in young adult mice. c, Proportion of dynamic sheaths engaging in growth three weeks before learning, during learning (one week), and in the two weeks after learning. d, Proportion of dynamic sheaths engaging in retraction before, during and after learning. Learning modulates sheath retraction (F2,14 = 6.76, p = 0.0088). During learning, more sheaths retract relative to untrained mice (p = 0.0095; Tukey’s HSD) and relative to prelearning values in trained mice (p = 0.016; Tukey’s HSD). Two weeks after learning, more sheaths retract relative to prelearning values in trained mice (p = 0.0358; Tukey’s HSD). e, In learning mice, there are fewer stable sheaths with 2 (p = 0.013, t(5.91)=3.49), 1 (p = 0.0089, t(2.47)=7.62), and 0 neighbors (p = 0.043, t(2.04)=4.55). f, In learning mice, there are more growing sheaths with 1 neighbor (p = 0.0071, t(6.00)=−4.00). g, In learning mice, there are more retracting sheaths with 2 (p = 0.024, t(4.00)=−3.62) and 1 neighbors (p = 0.0050, t(6.00)=−4.31). h, Nodes and gaps in myelin that shorten are not modulated by learning (REML with post hoc Tukey’s HSD). Bars and error bars represent mean ± s.e.m unless otherwise noted. For detailed statistics, see Supplementary Table 3.

Extended Data Fig. 5 Tracking virally labeled axons across time in vivo.

a, Maximum projection of a cFos+ neuron (AAv8-cFos-ER, red) and surrounding oligodendrocytes and myelin (in vivo, green) (top). Single slice at 159 microns below the pial surface, with white arrow identifying unmyelinated region of a cFos+ axon (middle). Single slice at 185 microns below the pial surface, with pink arrow identifying a myelinated region of a cFos+ axon (bottom). b, 3D projection of the cFos+ neuron and associated myelin sheaths from (a), with dashed lines corresponding to the single slices in (a). c, The same neuron as in (a), reconstructed using Simple Neurite Tracer. d, 3D projection of a longitudinally imaged axon with node lengthening and sheath retraction at the outset of the imaging experiment (0d) and at the final imaging timepoint (88d).

Extended Data Fig. 6 Biological correlates of computational modeling data.

a, Average pre-existing myelin sheath length per mouse (nmice = 4). b, Mean node length of lengthening nodes per mouse, measured across time in prelearning, learning, and postlearning stages (nmice = 3). c, Distribution of the number of consecutive lengthening nodes along a single axon. d-g, Modeled change in conduction speed as a function of the number of consecutively remodeled nodes at different sodium:leak conductance ratios (node with = 35 microns, pink exponential fit; Chi-square goodness of fit test). Conductance ratios were generated using either 0.4 S/cm2 (low) or 3.4 S/cm2 (high) for sodium conductance (gNa) and either 0.01 S/cm2 (low) or 0.08 S/cm2 (high) for leak conductance (gL). h, Length of unmyelinated axon segment that received new sheath addition following learning, measured from one sheath to the nearest neighboring sheath (nmice = 6; box plot, white bar = median, bounds of rectangle are Q1 and Q3; whiskers reflect minima and maxima). i, Proportion of gaps filled by 1, 2, or 3 sheaths following learning. j, Modeled change in conduction speed as a function of proportion of a 184 micron gap filled in by new myelin (pink exponential fit; Chi-square goodness of fit test). Conductance ratios as in d-g. Bars and error bars represent mean ± s.e.m unless otherwise noted.

Extended Data Fig. 7 Characterizing AAv8-cFos-ER virus.

a, Maximum projection of motor cortex injected with the cFos virus 1 day before learning (−1d), one week after learning (7d), three weeks after learning (22d), and 5 weeks after learning (42d). 4-OHT was injected 3 hours after the final day of learning (7d). b, Fit curve for modeled change in cFos+ cell gain (calculated as % of tamoxifen-independent cFos+ neurons, that is the number of neurons labeled at 60 days) in mice that learn and receive tamoxifen (pink), untrained mice (gray), and mice that learn and are injected with sunflower oil (red). Each dot represents proportion of labeled neurons per mouse at a given timepoint. Shading represents 95% confidence interval. c, Rate of cFos+ cell gain in mice that learn and receive tamoxifen (pink), untrained mice (grey), and mice that learn and are injected with sunflower oil (red). Mice that learn have a heightened rate of cFos+ cell gain in the two weeks following learning, while the percentage of cFos+ neurons did not change in untrained mice and oil injected mice over the course of the experiment. Lines and shading represent mean ± s.e.m. Light pink lines in background represent individual learning mice. d, Significantly more cFos+ cells appear in L2/3 relative to L1 following learning and injection of tamoxifen. e, The majority of learning-activated cells are putative L2/3 pyramidal neurons (determined by morphology). f, Distribution of traced axon lengths for cFos+ neurons (nmice = 5; box plot, white bar = median, bounds of rectangle are Q1 and Q3; whiskers reflect minima and maxima). Bars and error bars represent mean ± s.e.m unless otherwise noted. For detailed statistics, see Supplementary Table 3.

Supplementary information

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bacmeister, C.M., Huang, R., Osso, L.A. et al. Motor learning drives dynamic patterns of intermittent myelination on learning-activated axons. Nat Neurosci 25, 1300–1313 (2022). https://doi.org/10.1038/s41593-022-01169-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-022-01169-4

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing