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:

Unsupervised restoration of a complex learned behavior after large-scale neuronal perturbation

Abstract

Reliable execution of precise behaviors requires that brain circuits are resilient to variations in neuronal dynamics. Genetic perturbation of the majority of excitatory neurons in HVC, a brain region involved in song production, in adult songbirds with stereotypical songs triggered severe degradation of the song. The song fully recovered within 2 weeks, and substantial improvement occurred even when animals were prevented from singing during the recovery period, indicating that offline mechanisms enable recovery in an unsupervised manner. Song restoration was accompanied by increased excitatory synaptic input to neighboring, unmanipulated neurons in the same brain region. A model inspired by the behavioral and electrophysiological findings suggests that unsupervised single-cell and population-level homeostatic plasticity rules can support the functional restoration after large-scale disruption of networks that implement sequential dynamics. These observations suggest the existence of cellular and systems-level restorative mechanisms that ensure behavioral resilience.

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: Song degradation and recovery after selective large-scale perturbation of projection neurons.
Fig. 2: Dynamics of song degradation and recovery after large-scale perturbation of HVC projection neurons.
Fig. 3: Substantial song recovery without practice.
Fig. 4: Songs continue to recover rapidly during the postprevention period.
Fig. 5: Population-level homeostatic plasticity and recruitment of silent neurons in a network model contribute to the recovery of sequential activity.
Fig. 6: Saltatory recovery of syllable duration.

Similar content being viewed by others

Data availability

All derived data in this study are included in this article. Raw datasets are publicly available online at https://doi.org/10.22002/dvhsa-h5s72 (ref. 72) or by contacting the corresponding authors. Source data are provided with this paper.

Code availability

Custom codes associated with this study are publicly available (for behavior analysis, https://doi.org/10.5281/zenodo.10823142 (ref. 73); for modeling, code is available on GitHub at https://github.com/davidgbe/unsupervised_restoration_modeling (ref. 74) and on Zenodo at https://doi.org/10.5281/zenodo.10823218 (ref. 75)).

References

  1. Masset, P., Qin, S. & Zavatone-Veth, J. A. Drifting neuronal representations: bug or feature? Biol. Cybern. 116, 253–266 (2022).

    Article  PubMed  Google Scholar 

  2. Mongillo, G., Rumpel, S. & Loewenstein, Y. Intrinsic volatility of synaptic connections—a challenge to the synaptic trace theory of memory. Curr. Opin. Neurobiol. 46, 7–13 (2017).

    Article  CAS  PubMed  Google Scholar 

  3. Nottebohm, F., Stokes, T. M. & Leonard, C. M. Central control of song in the canary, Serinus canarius. J. Comp. Neurol. 165, 457–486 (1976).

    Article  CAS  PubMed  Google Scholar 

  4. Nottebohm, F., Kelley, D. B. & Paton, J. A. Connections of vocal control nuclei in the canary telencephalon. J. Comp. Neurol. 207, 344–357 (1982).

    Article  CAS  PubMed  Google Scholar 

  5. Hahnloser, R. H. R., Kozhevnikov, A. A. & Fee, M. S. An ultra-sparse code underlies the generation of neural sequences in a songbird. Nature 419, 65–70 (2002).

    Article  CAS  PubMed  Google Scholar 

  6. Long, M. A. & Fee, M. S. Using temperature to analyse temporal dynamics in the songbird motor pathway. Nature 456, 189–194 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Hamaguchi, K., Tanaka, M. & Mooney, R. A distributed recurrent network contributes to temporally precise vocalizations. Neuron 91, 680–693 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Coleman, M. J. & Vu, E. T. Recovery of impaired songs following unilateral but not bilateral lesions of nucleus uvaeformis of adult zebra finches. J. Neurobiol. 63, 70–89 (2005).

    Article  PubMed  Google Scholar 

  9. Thompson, J. A., Wu, W., Bertram, R. & Johnson, F. Auditory-dependent vocal recovery in adult male zebra finches is facilitated by lesion of a forebrain pathway that includes the basal ganglia. J. Neurosci. 27, 12308–12320 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Poole, B., Markowitz, J. E. & Gardner, T. J. The song must go on: resilience of the songbird vocal motor pathway. PLoS ONE 7, e38173 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Otchy, T. M. et al. Acute off-target effects of neural circuit manipulations. Nature 528, 358–363 (2015).

    Article  CAS  PubMed  Google Scholar 

  12. Long, M. A., Jin, D. Z. & Fee, M. S. Support for a synaptic chain model of neuronal sequence generation. Nature 468, 394–399 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Ren, D. et al. A prokaryotic voltage-gated sodium channel. Science 294, 2372–2375 (2001).

    Article  CAS  PubMed  Google Scholar 

  14. Liberti, W. A. et al. Unstable neurons underlie a stable learned behavior. Nat. Neurosci. 19, 1665–1671 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Markowitz, J. E. et al. Mesoscopic patterns of neural activity support songbird cortical sequences. PLoS Biol. 13, e1002158 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Sim, S., Antolin, S., Lin, C.-W., Lin, Y. & Lois, C. Increased cell-intrinsic excitability induces synaptic changes in new neurons in the adult dentate gyrus that require Npas4. J. Neurosci. 33, 7928–7940 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Lin, C.-W. et al. Genetically increased cell-intrinsic excitability enhances neuronal integration into adult brain circuits. Neuron 65, 32–39 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Xue, M., Atallah, B. V. & Scanziani, M. Equalizing excitation–inhibition ratios across visual cortical neurons. Nature 511, 596–600 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. McInnes, L., Healy, J. & Melville, J. UMAP: uniform manifold approximation and projection for dimension reduction. Preprint at arXiv https://doi.org/10.48550/arXiv.1802.03426 (2018).

  20. Goffinet, J., Brudner, S., Mooney, R. & Pearson, J. Low-dimensional learned feature spaces quantify individual and group differences in vocal repertoires. eLife 10, e67855 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Sainburg, T., Thielk, M. & Gentner, T. Q. Finding, visualizing, and quantifying latent structure across diverse animal vocal repertoires. PLoS Comput. Biol. 16, e1008228 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Turrigiano, G. G. & Nelson, S. B. Homeostatic plasticity in the developing nervous system. Nat. Rev. Neurosci. 5, 97–107 (2004).

    Article  CAS  PubMed  Google Scholar 

  23. Sweeney, S. T., Broadie, K., Keane, J., Niemann, H. & O’Kane, C. J. Targeted expression of tetanus toxin light chain in Drosophila specifically eliminates synaptic transmission and causes behavioral defects. Neuron 14, 341–351 (1995).

    Article  CAS  PubMed  Google Scholar 

  24. Tchernichovski, O., Nottebohm, F., Ho, C. E., Pesaran, B. & Mitra, P. P. A procedure for an automated measurement of song similarity. Anim. Behav. 59, 1167–1176 (2000).

    Article  CAS  PubMed  Google Scholar 

  25. Tchernichovski, O., Mitra, P. P., Lints, T. & Nottebohm, F. Dynamics of the vocal imitation process: how a zebra finch learns its song. Science 291, 2564–2569 (2001).

    Article  CAS  PubMed  Google Scholar 

  26. Jun, J. K. & Jin, D. Z. Development of neural circuitry for precise temporal sequences through spontaneous activity, axon remodeling, and synaptic plasticity. PLoS ONE 2, e723 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  27. 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 

  28. Fiete, I. R., Senn, W., Wang, C. Z. H. & Hahnloser, R. H. R. Spike-time-dependent plasticity and heterosynaptic competition organize networks to produce long scale-free sequences of neural activity. Neuron 65, 563–576 (2010).

    Article  CAS  PubMed  Google Scholar 

  29. Veliz-Cuba, A., Shouval, H. Z., Josić, K. & Kilpatrick, Z. P. Networks that learn the precise timing of event sequences. J. Comput. Neurosci. 39, 235–254 (2015).

    Article  PubMed  Google Scholar 

  30. Murray, J. M. & Escola, G. S. Learning multiple variable-speed sequences in striatum via cortical tutoring. eLife 6, e26084 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  31. Kim, J. & Tsien, R. W. Synapse-specific adaptations to inactivity in hippocampal circuits achieve homeostatic gain control while dampening network reverberation. Neuron 58, 925–937 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Jarvis, E. D., Scharff, C., Grossman, M. R., Ramos, J. A. & Nottebohm, F. For whom the bird sings: context-dependent gene expression. Neuron 21, 775–788 (1998).

    Article  CAS  PubMed  Google Scholar 

  33. Kozhevnikov, A. A. & Fee, M. S. Singing-related activity of identified HVC neurons in the zebra finch. J. Neurophysiol. 97, 4271–4283 (2007).

    Article  PubMed  Google Scholar 

  34. Hirase, H., Leinekugel, X., Czurkó, A., Csicsvari, J. & Buzsáki, G. Firing rates of hippocampal neurons are preserved during subsequent sleep episodes and modified by novel awake experience. Proc. Natl Acad. Sci. USA 98, 9386–9390 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Slomowitz, E. et al. Interplay between population firing stability and single neuron dynamics in hippocampal networks. eLife 4, e04378 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Trouche, S. et al. Recoding a cocaine-place memory engram to a neutral engram in the hippocampus. Nat. Neurosci. 19, 564–567 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Hartman, K. N., Pal, S. K., Burrone, J. & Murthy, V. N. Activity-dependent regulation of inhibitory synaptic transmission in hippocampal neurons. Nat. Neurosci. 9, 642–649 (2006).

    Article  CAS  PubMed  Google Scholar 

  38. Beattie, E. C. et al. Control of synaptic strength by glial TNF-alpha. Science 295, 2282–2285 (2002).

    Article  CAS  PubMed  Google Scholar 

  39. Schinder, A. F. & Poo, M. The neurotrophin hypothesis for synaptic plasticity. Trends Neurosci. 23, 639–645 (2000).

    Article  CAS  PubMed  Google Scholar 

  40. Wang, Y. et al. Astrocyte-secreted IL-33 mediates homeostatic synaptic plasticity in the adult hippocampus. Proc. Natl Acad. Sci. USA 118, e2020810118 (2021).

    Article  CAS  PubMed  Google Scholar 

  41. Stellwagen, D. & Malenka, R. C. Synaptic scaling mediated by glial TNF-alpha. Nature 440, 1054–1059 (2006).

    Article  CAS  PubMed  Google Scholar 

  42. Scharff, C., Kirn, J. R., Grossman, M., Macklis, J. D. & Nottebohm, F. Targeted neuronal death affects neuronal replacement and vocal behavior in adult songbirds. Neuron 25, 481–492 (2000).

    Article  CAS  PubMed  Google Scholar 

  43. Nicola, W. & Clopath, C. Supervised learning in spiking neural networks with FORCE training. Nat. Commun. 8, 2208 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Adam, I. et al. Daily vocal exercise is necessary for peak performance singing in a songbird. Nat. Commun. 14, 7787 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Katlowitz, K. A., Picardo, M. A. & Long, M. A. Stable sequential activity underlying the maintenance of a precisely executed skilled behavior. Neuron 98, 1133–1140.e3 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Schoonover, C. E., Ohashi, S. N., Axel, R. & Fink, A. J. P. Representational drift in primary olfactory cortex. Nature 594, 541–546 (2021).

    Article  CAS  PubMed  Google Scholar 

  47. Gonzalez, W. G., Zhang, H., Harutyunyan, A. & Lois, C. Persistence of neuronal representations through time and damage in the hippocampus. Science 365, 821–825 (2019).

    Article  CAS  PubMed  Google Scholar 

  48. Driscoll, L. N., Pettit, N. L., Minderer, M., Chettih, S. N. & Harvey, C. D. Dynamic reorganization of neuronal activity patterns in parietal cortex. Cell 170, 986–999.e16 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Marks, T. D. & Goard, M. J. Stimulus-dependent representational drift in primary visual cortex. Nat. Commun. 12, 5169 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Ziv, Y. et al. Long-term dynamics of CA1 hippocampal place codes. Nat. Neurosci. 16, 264–266 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Duffy, A., Abe, E., Perkel, D. J. & Fairhall, A. L. Variation in sequence dynamics improves maintenance of stereotyped behavior in an example from bird song. Proc. Natl Acad. Sci. USA 116, 9592–9597 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Rule, M. E. & O’Leary, T. Self-healing codes: how stable neural populations can track continually reconfiguring neural representations. Proc. Natl Acad. Sci. USA 119, e2106692119 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Dave, A. S. & Margoliash, D. Song replay during sleep and computational rules for sensorimotor vocal learning. Science 290, 812–816 (2000).

    Article  CAS  PubMed  Google Scholar 

  54. Kao, M. H., Doupe, A. J. & Brainard, M. S. Contributions of an avian basal ganglia–forebrain circuit to real-time modulation of song. Nature 433, 638–643 (2005).

    Article  CAS  PubMed  Google Scholar 

  55. Brainard, M. S. & Doupe, A. J. Interruption of a basal ganglia–forebrain circuit prevents plasticity of learned vocalizations. Nature 404, 762–766 (2000).

    Article  CAS  PubMed  Google Scholar 

  56. Williams, H. & Mehta, N. Changes in adult zebra finch song require a forebrain nucleus that is not necessary for song production. J. Neurobiol. 39, 14–28 (1999).

    Article  CAS  PubMed  Google Scholar 

  57. Vu, E. T., Schmidt, M. F. & Mazurek, M. E. Interhemispheric coordination of premotor neural activity during singing in adult zebra finches. J. Neurosci. 18, 9088–9098 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Wang, C. Z. H., Herbst, J. A., Keller, G. B. & Hahnloser, R. H. R. Rapid interhemispheric switching during vocal production in a songbird. PLoS Biol. 6, e250 (2008).

    Article  PubMed  Google Scholar 

  59. Lois, C., Hong, E. J., Pease, S., Brown, E. J. & Baltimore, D. Germline transmission and tissue-specific expression of transgenes delivered by lentiviral vectors. Science 295, 868–872 (2002).

    Article  CAS  PubMed  Google Scholar 

  60. Kollmorgen, S., Hahnloser, R. H. R. & Mante, V. Nearest neighbours reveal fast and slow components of motor learning. Nature 577, 526–530 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Meehan, C., Ebrahimian, J., Moore, W. & Meehan, S. Uniform manifold approximation and projection (UMAP). MATLAB Central File Exchange https://www.mathworks.com/matlabcentral/fileexchange/71902 (2022).

  62. Mooney, R. Different subthreshold mechanisms underlie song selectivity in identified HVC neurons of the zebra finch. J. Neurosci. 20, 5420–5436 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Garst-Orozco, J., Babadi, B. & Ölveczky, B. P. A neural circuit mechanism for regulating vocal variability during song learning in zebra finches. eLife 3, e03697 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Spiro, J. E., Dalva, M. B. & Mooney, R. Long-range inhibition within the zebra finch song nucleus RA can coordinate the firing of multiple projection neurons. J. Neurophysiol. 81, 3007–3020 (1999).

    Article  CAS  PubMed  Google Scholar 

  65. Tupikov, Y. & Jin, D. Z. Addition of new neurons and the emergence of a local neural circuit for precise timing. PLoS Comput. Biol. 17, e1008824 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Mooney, R. & Prather, J. F. The HVC microcircuit: the synaptic basis for interactions between song motor and vocal plasticity pathways. J. Neurosci. 25, 1952–1964 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Kornfeld, J. et al. EM connectomics reveals axonal target variation in a sequence-generating network. eLife 6, e24364 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  68. Thiagarajan, T. C., Piedras-Renteria, E. S. & Tsien, R. W. Alpha- and betaCaMKII. Inverse regulation by neuronal activity and opposing effects on synaptic strength. Neuron 36, 1103–1114 (2002).

    Article  CAS  PubMed  Google Scholar 

  69. Thiagarajan, T. C., Lindskog, M. & Tsien, R. W. Adaptation to synaptic inactivity in hippocampal neurons. Neuron 47, 725–737 (2005).

    Article  CAS  PubMed  Google Scholar 

  70. Han, E. B. & Stevens, C. F. Development regulates a switch between post- and presynaptic strengthening in response to activity deprivation. Proc. Natl Acad. Sci. USA 106, 10817–10822 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Pfister, J.-P. & Gerstner, W. Triplets of spikes in a model of spike timing-dependent plasticity. J. Neurosci. 26, 9673–9682 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Wang, et al. Datasets for “Unsupervised restoration of a complex learned behavior after large-scale neuronal perturbation”. Caltech Library https://doi.org/10.22002/dvhsa-h5s72 (2024).

  73. Duffy, A. Behavioral analysis. Zenodo https://doi.org/10.5281/zenodo.10823142 (2024).

  74. unsupervised_restoration_modeling. GitHub https://github.com/davidgbe/unsupervised_restoration_modeling (2024).

  75. Bell, D. Unsupervised restoration of a complex learned behavior after large-scale neuronal perturbation - modeling. Zenodo https://doi.org/10.5281/zenodo.10823218 (2024).

Download references

Acknowledgements

This work was supported by National Institutes of Health grant R01 NS104925-01 (C.L. and A.L.F.). We thank R. Pang for contributing to the code used to run simulations. We thank F. Lagzi for valuable discussions. This work was facilitated through the use of advanced computational, storage and networking infrastructure provided by the Hyak supercomputer system and funded by the Student Technology Fee at the University of Washington.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed to the conceptualization, methodology, and writing of the paper. Experimental investigations were undertaken by B.W., Z.T., S.W., T.A.F.V. and C.L. Computational investigations were carried out by A.D. and D.G.B. Visualization was performed by B.W., A.D. and D.G.B. Funding acquisition and supervision was done by A.L.F. and C.L.

Corresponding authors

Correspondence to Bo Wang or Carlos Lois.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Neuroscience thanks Michael Long 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 Specific infection of HVC projection neurons by LVs.

(a) Confocal image of a brain slice from a bird injected with LVs, showing the expression of the transgene (tagged with GFP) in HVC neurons. HVC(RA) neurons are labeled by a fluorescent retrograde tracer (cholera toxin b - alexa 555) injected into RA; (b) Confocal images of a brain slice showing that LV selectively target projection neurons, where only 1/1000 of the cells labeled by LV transgene (tagged with GFP, seen in green) overlap the immunofluorescent signal of pooled antibodies against some of the standard markers of inhibitory neurons (PV, parvalbumin/CB, calbindin/CR, calretinin, seen in red and blue), N = 2 animals; (c) Example of serial sagittal slices of HVC, marked by their distance relative to the most lateral side of HVC, showing the efficient labeling of HVC neurons by LV, except for the most medial posterior corner (as demonstrated by the +1300 and +1400 μm images); (d) Confocal images showing the expression of the viral-delivered transgene (tagged with GFP) in the majority of projection neurons (retrogradely labeled with fluorescent tracer, FluoroGold/cholera toxin b, injected into RA and X, respectively); (e) Histogram indicating the percentage of GFP-labeled cells among all projection neurons in HVC. Each pair of bars is generated from one animal. N = 3 animals. Error bars indicate 95% confidence interval. Numbers in each bar are number of identified GFP positive cells divided by total number of cells counted.

Extended Data Fig. 2 Expression of NaChBac in HVC(RA) neurons.

(a) Schematic drawing showing whole-cell patch clamp recordings made in HVC(RA) neurons infected with LV-NaChBac or naive control; (b) Example traces of whole-cell NaChBac currents evoked by depolarizing voltage steps (from −80 to +30 mV. Increment, 10 mV) recorded from HVC(RA) neurons; (c) I-V curve of the NaChBac current, illustrating the peak amplitude of whole-cell current at different step voltages; (d) Comparison of the maximal peak amplitude of whole-cell NaChBac currents recorded at 5 dpi (3.6 ± 0.6 nA) vs. 21–28 dpi (3.0 ± 0.3 nA). Student’s t-test. Sample size is shared between panels b and c, N = 10 cells / 3 animals (control), 15 / 2 (5 dpi), and 19 / 3 (21–28 dpi); (e) Example of current traces with mIPSC events recorded in HVC(RA) neurons expressing NaChBac at different times after injection. The bar called ‘Degraded’ illustrates currents at 5 dpi. The bar called ‘Recovered’ illustrates currents at 25–35 dpi; (f) Group data of the frequency and amplitude of mIPSCs in HVC(RA) NaChBac+ cells. mIPSC frequency: Control, 2.7 ± 0.2 s−1, N = 23 / 4; Degraded, 6.1 ± 0.7 s−1, N = 17 / 3; Recovered, 7.2 ± 0.7 s−1, N = 16 / 5; (g) Example of current traces with mEPSC events recorded in HVC(RA) neurons expressing NaChBac at different time points; (h) Group data of the frequency and amplitude of mEPSCs in HVC(RA) NaChBac+ cells. mEPSC frequency: Control, 9.7 ± 1.6 min−1, N = 22 / 4; Degraded, 6.8 ± 0.9 min−1, N = 14 / 5; Recovered, 5.9 ± 0.9 min−1, N = 17 / 4. mEPSC amplitude: Control, 17.2 ± 0.9 pA; Degraded, 18.5 ± 0.9 pA; Recovered, 16.1 ± 0.5 pA. One-way ANOVA & student’s t-test; (i) Confocal images of an HVC slice stained with antibodies against GFP and EGR-1, an immediate early gene whose expression is induced by singing related activity. Scale bar, 25 μm. Scatter plot in the bottom right represents the percentage of GFP-positive cells that are co-expressing EGR-1. Each dot represents data from one bird. N = 3 (4 dpi) and 4 (28 dpi) animals. Student’s t-test. Error bars represent SEM.

Source data

Extended Data Fig. 3 Expressing TeNT in HVC(RA) cells blocked their synaptic output.

(a) Example traces from a dual patch clamp recording in control animals, showing that when action potentials were evoked in HVC(RA) neurons (upper), excitatory postsynaptic currents (EPSCs) can be reliably detected in the connected interneuron (below); (b) Example traces similar to those in panel a, showing that no EPSC could be detected in the interneuron when the HVC(RA) neuron was expressing TeNT; (c) Summary of dual patch clamp results. The amplitude of EPSCs are shown in the scatter plot. The block of synaptic transmission by TeNT expression is long-lasting, as it could be detected even >3 weeks after viral injection, after the song had already recovered; (d) Confocal image showing axon bundles from HVC and their terminal branches within RA labeled by LV-TeNT (tagged by GFP); (e) To evoke post-synaptic responses in RA neurons, a bipolar stimulation electrode was placed in the vicinity of the axon bundles going into RA. Under DIC microscopy, RA is clearly distinguishable by strong phase contrast against the surrounding region. Scale bar, 100 μm; (f) Whole-cell patch clamp recordings were made in RA neurons. Example membrane potential traces in response to depolarizing and hyperpolarizing current steps recorded from one projection neuron in RA (RA(PN)); (g) In RA(PN) neurons, inward post-synaptic current transients were evoked by briefly stimulating axon bundles from HVC with varying intensity, measured by amplitude of stimulus current. Compared with naïve animals (black traces on top), the responses were much smaller in RA(PN) neurons recorded in birds with LV-TeNT in HVC (green traces below); (h) Summarized data of the synaptic responses in RA(PN) neurons. Stimulus – response curves were composed with data obtained at three different time points. The central marks of box plots indicate the medians, the bottom and top edges indicate the 1st and 3rd quartiles, and the whiskers indicate the most extreme data points that are still within the 1.5 times of the inter-quartile-region beyond the lower and upper quartiles. Outliers are individually marked. N = 11 cells / 3 animals (Control), 15 / 3 (2 dpi), and 12 / 1 (21–28 dpi).

Source data

Extended Data Fig. 4 Songs degradation and recovery was not due to mechanical lesion or inflammation.

(a) Example spectrograms of songs from a bird injected with LV-NaChBac(EtoK), a dead-pore mutant of NaChBac; (b) Distribution of syllable durations per day of the same bird shown in panel a; (c) Plots of scaled acoustic distance to original syllables. Data was from the same bird as in panels a and b. The insets are UMAP visualizations of songs at selected time points. Dashed lines are generated from the syllables of the bird injected with LV-NaChBac shown in Fig. 1d, and for comparison here they were not normalized to maximum. Experiments with LV-NaChBac(EtoK) yielded consistent results, N = 4 animals; (d) Example spectrograms of songs from a bird injected with half of the volume of LV-NaChBac used for animals shown in Fig. 1; (e) Distribution of syllable durations per day of the same bird shown in panel d; (f) Plots of scaled acoustic distance to original syllables. Data was from the same bird as in panels d and e. Each line/color represents one syllable. The insets are UMAP visualizations of songs at selected time points. Experiments with reduced volume of LV-NaChBac were repeated, N = 2.

Extended Data Fig. 5 Recovery of fine intra-syllable structure without practice.

(a-d) Individual curves of time-varying acoustic features, including mean frequency (a), entropy (b), pitch (c), and goodness of pitch (d), generated from the same syllable as shown in Fig. 3g&h. Each black curve represents one rendition of the syllable, and the red curve represents the mean. The population of renditions shown are the 2.5% of renditions closest to the original syllable cluster in a given day. (e-j) More example syllables to show intra-syllable acoustic structures on the first day after song prevention was lifted. Four time points were selected for comparison, pre-perturbation, pre-prevention, first day postprevention, and when songs were fully recovered. Spectrograms (on the left in each panel) and averaged time-varying acoustic features (right) of each corresponding syllable were plotted in the same way as Fig. 3e–h. All spectrograms share the same time scale. Note that, on the first day after song-prevention (postprevention), clear syllable renditions were sung and the intra-syllable timing of acoustic feature fluctuations were very similar to pre-perturbation syllables, though differences in acoustic feature magnitudes were still evident and the rendition-to-rendition variability was higher. Shading represents standard error of mean.

Extended Data Fig. 6 Recovery of syllable durations without practice.

(a) The distribution of durations of one example syllable, same as the one shown in Fig. 3g&h, before perturbation and right after song prevention (bin size, 10 ms). Note that they both peak at the same value. Two syllable-selection methods were used (see the Methods), with or without dynamic time warping, in order to test whether the k-nearest neighbor measure was missing dilated or compressed syllable renditions. The two syllable-selection methods found highly similar duration distributions; (b) Comparison of the mode of syllable durations before perturbation and post-prevention. Most syllables recovered their duration modes with high precision without song practice. Dashed lines indicate second peak of dual peak distribution; (c) Comparison of the standard deviation of duration distributions demonstrates that the variability of syllable durations was still elevated the first day post prevention. Comparable results were obtained with either type of syllable-selection method. In a-c, we performed this analysis on 2/3 birds in the song prevention experiment. In the third bird, the motif was not identifiable the first day post-song prevention, making it not feasible to construct individual syllable duration distributions on this day. One potential reason the song recovery trajectory differed is that this bird’s song was abnormally variable before the viral perturbation and song prevention experiment. We include this bird (V449) in the data and analyses presented in Fig. 4b–e where the pre-perturbation variability and the post-prevention lag in recovery are clear. N = 8 syllables / 2 animals for both panel b and c.

Extended Data Fig. 7 Unperturbed cells did not change their intrinsic excitability or inhibitory synaptic inputs.

(a) (Left) Whole-cell patch clamp recordings were made in GFP-negative HVC(RA) neurons in naive control birds or birds injected with LV-TeNT. (Right) Membrane potential and firing pattern of HVC(RA) neurons in response to current steps; (b) Group data showing that no significant difference was found in the resting membrane potential (−75.2 ± 0.8 vs. −74.8 ± 0.8 mV), input resistance (414.8 ± 30.0 vs. 416.2 ± 33.1 MOhm), or initial F-I slope (437.9 ± 35.5 vs. 406.2 ± 37.3 Hz/A) between neurons in naive control (N = 22 cells / 3 animals) and in birds with LV-TeNT for more than 25 days (N = 30 / 4). Student’s t-test; (c) F-I curves obtained from HVC(RA) neurons showed no difference between control or animals recovered from LV-TeNT. Sample size is the same as that in panel b; (d) Group data of mIPSC recorded in HVC(RA) neurons, showing no significant difference at any time during the experiment. ‘Degraded’ indicates the time when the song was degraded, at 5 dpi. ‘Recovered’ indicates the time after the song had fully recovered, at 25 dpi. mIPSC frequency: Control, 2.7 ± 0.2 s−1, N = 23 cells / 4 animals; ‘Degraded’ GFP-negative cell, 2.6 ± 0.4 s−1, N = 16 / 3; ‘Degraded’ GFP-positive, 2.3 ± 0.4 s−1, N = 16 / 3; ‘Recovered’ GFP-, 2.6 ± 0.5 s−1, N = 9 / 3; ‘Recovered’ GFP+, 3.1 ± 0.8 s−1, N = 9 / 3. mIPSC amplitude: Control, 37.9 ± 2.0 pA; ‘Degraded’ GFP-negative cell, 40.6 ± 1.7 pA; ‘Degraded’ GFP-positive cell, 43.2 ± 1.6 pA; ‘Recovered’ GFP- negative cell, 39.0 ± 3.7 pA; ‘Recovered’ GFP- positive cell, 41.4 ± 5.4 pA. One-way ANOVA. Error bars represent SEM.

Source data

Extended Data Fig. 8 Synaptic changes in unmanipulated neurons only occurred in the injected hemisphere.

(a) Example song spectrograms of a bird with LV-TeNT in one side of the two HVCs; (b) Acoustic distance to each original syllable when LV-TeNT was injected bilaterally (left) or unilaterally (right). Curves in the same color are from the same bird. Bilateral trajectories are similar to those shown in Fig. 2f but without normalization; (c) Comparison of the peak distortion, measured by the maximal distance from each original syllable, between unilateral or bilateral LV-TeNT perturbations. Each column represents syllables from the same bird. N = 24 syllables (nested) / 4 animals for Bilateral and 20 (nested) / 3 for Unilateral. Nested two-way ANOVA; (d) After unilateral LV-TeNT perturbation, the day when each syllable reached peak distortion compared with bilateral LV-TeNT (triangles. On average, 2.75 ± 0.46 dpi. Nested two-way ANOVA) and the day when each syllable achieved more than 90% recovery (round dots. on average, 17.82 ± 2.49 dpi, later than that for bilateral LV-TeNT, 13.71 ± 0.72 dpi, nested two-way ANOVA). Sample size is the same as that in panel c; (e) Whole-cell recordings were made in GFP-negative HVC(RA) (‘unmanipulated’) neurons and interneurons in the injected and unperturbed hemispheres of birds with unilateral LV-TeNT injection; (f) Group data of mEPSC in naive control birds and birds with unilateral LV-TeNT. (Middle) Cumulative curve of inter-event intervals of mEPSCs. Dashed line represents data from GFP- neurons from animals with bilateral LV-TeNT injection for comparison and was adapted from Fig. 5c. mEPSC frequency: Naive control, 9.7 ± 1.6 min−1, N = 23 cells / 4 animals; Contralateral (uninjected HVC), 9.8 ± 2.3 min−1, N = 16 / 4; Ipsilateral (injected), 22.8 ± 3.3 min−1, N = 18 / 3. mEPSC amplitude: Naive control, 17.2 ± 0.9 pA; Contralateral, 18.2 ± 0.9 pA; Ipsilateral, 22.3 ± 1.0 pA. One-way ANOVA followed by t-test (without adjustment); (g) Group data of mEPSC recorded in interneurons. (Left) The frequency of mEPSC in interneurons decreased after virus injection, but eventually recovered to a level comparable to that of controls. (Middle) Cumulative curve of inter-event intervals of mEPSCs. mEPSC frequency: Naive control, 1.47 ± 0.14 s−1, N = 19 / 2; Degraded contralateral, 1.56 ± 0.16 s−1, N = 13 / 3; Degraded ipsilateral, 0.63 ± 0.14 s−1; N = 20 / 3; Recovered contralateral, 1.34 ± 0.19 s−1, N = 18 / 4; Recovered ipsilateral, 1.45 ± 0.15 s−1, N = 20 / 4. mEPSC amplitude: Naive control, 38.3 ± 1.2 pA; Degraded contralateral, 38.4 ± 2.6 pA; Degraded ipsilateral, 35.5 ± 1.9 pA; Recovered contralateral, 37.3 ± 2.1 pA; Recovered ipsilateral, 33.2 ± 2.0 pA.One-way ANOVA & student’s t-test. Error bars represent SEM.

Source data

Extended Data Fig. 9 Recovery of modeled sequences with STDP.

(a) Schematic showing triplet STDP implemented both in E→E and E→I synapses. Potentiation is mediated by triplets of spikes; depression is dictated by a pairwise rule. See Methods for details; (b) Spike raster plots showing the sequential dynamics generated by HVC neurons before and after perturbation with only STDP and downward firing rate homeostasis implemented (see Eq. 2 in Methods). Colors represent the firing timing of each cell before perturbation. Note that the sequence regenerates serially and all cells recapture their original firing timing; (c-e) Percentage of syllables completed prior to (pre-perturbation), just after (perturbed), and 3000 renditions after (recovered) perturbation for networks with (c) STDP alone, (d) STDP plus Poisson inputs, (e) or firing rate homeostatic plasticity, for 30% to 70% perturbations in increments of 10%. Note that networks with firing rate homeostatic mechanisms possessed larger recoverable regimes.

Extended Data Fig. 10 Recovery of modeled sequences with population-level homeostatic plasticity.

(a-d) are generated from networks with STDP and a downward, single-cell, firing rate homeostasis rule implemented; (e-h) are generated from networks with STDP and 2-sided, single-cell, firing rate homeostasis; (i-l) are generated from networks with silent cells and population-level homeostasis added to rules in (a-d); (a,e,i) Percentage of modeled syllables that successfully completed following >80% of activations in three time windows (before perturbation, perturbed and recovered) in response to different degrees of perturbation (10, 25, and 50%), when different sets of plasticity rules were implemented. Each point represents a network. Lines represent averages over all networks (N = 15); (b,f,j) The normalized total firing activity of all functional HVC(RA) (left) or interneurons (right) plotted against renditions. Solid curves show the average of multiple networks (N = 15) and shading represents SEM. RPP, renditions post perturbation; (c,g,k) Normalized total excitatory synaptic input into each unmanipulated HVC(RA) neuron after perturbations. Note that with population-level plasticity and silent neurons, an increase in E → E weight over its pre-perturbation value was introduced; (d,h,l) Distributions of single-cell firing rates per rendition for HVC(RA) (top) and interneurons (bottom) before and following 50% perturbation. Pre-perturbation distributions in black or gray. Lines represent mean and shadings represent standard error of mean, N = 15.

Supplementary information

Supplementary Information

Supplementary Figs. 1–7.

Reporting Summary

Supplementary Audio 1

Example song of bird V287 before perturbation. Spectrogram is shown in Fig. 1d, preperturbation.

Supplementary Audio 2

Example song of bird V287 with LV-NaChBac. Spectrogram is shown in Fig. 1d, at 6 dpi.

Supplementary Audio 3

Example song of bird V287 after recovery. Spectrogram is shown in Fig. 1d, at 25 dpi.

Supplementary Audio 4

Example song of bird V405 before perturbation. Spectrogram is shown in Fig. 1f, preperturbation.

Supplementary Audio 5

Example song of bird V405 with LV-TeNT. Spectrogram is shown in Fig. 1f, at 4 dpi.

Supplementary Audio 6

Example song of bird V405 after recovery. Spectrogram is shown in Fig. 1f, at 25 dpi.

Source data

Source Data Fig. 2

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 2/Table 2

Statistical source data.

Source Data Extended Data Fig. 3/Table 3

Statistical source data.

Source Data Extended Data Fig. 7/Table 7

Statistical source data.

Source Data Extended Data Fig. 8/Table 8

Statistical source data.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) 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

Wang, B., Torok, Z., Duffy, A. et al. Unsupervised restoration of a complex learned behavior after large-scale neuronal perturbation. Nat Neurosci (2024). https://doi.org/10.1038/s41593-024-01630-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1038/s41593-024-01630-6

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