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:

Delayed motor learning in a 16p11.2 deletion mouse model of autism is rescued by locus coeruleus activation

Abstract

Children with autism spectrum disorder often exhibit delays in achieving motor developmental milestones such as crawling, walking and speech articulation. However, little is known about the neural mechanisms underlying motor-related deficits. Here, we reveal that mice with a syntenic deletion of the chromosome 16p11.2, a common copy number variation associated with autism spectrum disorder, also exhibit delayed motor learning without showing gross motor deficits. Using in vivo two-photon imaging in awake mice, we find that layer 2/3 excitatory neurons in the motor cortex of adult male 16p11.2-deletion mice show abnormally high activity during the initial phase of learning, and the process of learning-induced spine reorganization is prolonged. Pharmacogenetic activation of locus coeruleus noradrenergic neurons was sufficient to rescue the circuit deficits and the delayed motor learning in these mice. Our results unveil an unanticipated role of noradrenergic neuromodulation in improving the delayed motor learning in 16p11.2-deletion male mice.

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: Delayed motor learning in 16p11.2+/− mice.
Fig. 2: Abnormally high ensemble activity during the initial phase of motor learning in 16p11.2+/− mice.
Fig. 3: Delayed spine pruning in 16p11.2+/− mice.
Fig. 4: Fewer LC-NA neurons are activated in 16p11.2+/− mice during motor learning.
Fig. 5: Pharmacogenetic activation of LC-NA neurons in 16p11.2+/−::DBH-CreERT mice lowered the abnormally high ensemble activity and reversed the delayed spine pruning.
Fig. 6: Pharmacogenetic activation of LC-NA neurons in 16p11.2+/−::DBH-CreERT mice rescued the delayed motor learning.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author upon request.

Code availability

The codes used for analyzing the data in the current study are available from the corresponding author upon request.

References

  1. Baio, J. et al. Prevalence of autism spectrum disorder among children aged 8 years—Autism and Developmental Disabilities Monitoring Network, 11 sites, United States, 2014. MMWR Surveill. Summ. 67, 1–23 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  2. Provost, B., Lopez, B. R. & Heimerl, S. A comparison of motor delays in young children: autism spectrum disorder, developmental delay, and developmental concerns. J. Autism Dev. Disord. 37, 321–328 (2007).

    Article  PubMed  Google Scholar 

  3. Harris, S. R. Early motor delays as diagnostic clues in autism spectrum disorder. Eur. J. Pediatr. 176, 1259–1262 (2017).

    Article  PubMed  Google Scholar 

  4. Weiss, L. A. et al. Association between microdeletion and microduplication at 16p11.2 and autism. N. Engl. J. Med. 358, 667–675 (2008).

    Article  CAS  PubMed  Google Scholar 

  5. Hippolyte, L. et al. The number of genomic copies at the 16p11.2 locus modulates language, verbal memory, and inhibition. Biol. Psychiatry 80, 129–139 (2016).

    Article  CAS  PubMed  Google Scholar 

  6. Demopoulos, C. et al. Abnormal speech motor control in individuals with 16p11.2 deletions. Sci. Rep. 8, 1274 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  7. Yadav, S. et al. TAOK2 kinase mediates PSD95 stability and dendritic spine maturation through Septin7 phosphorylation. Neuron 93, 379–393 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Ip, J. P. K. et al. Major vault protein, a candidate gene in 16p11.2 microdeletion syndrome, is required for the homeostatic regulation of visual cortical plasticity. J. Neurosci. 38, 3890–3900 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Portmann, T. et al. Behavioral abnormalities and circuit defects in the basal ganglia of a mouse model of 16p11.2 deletion syndrome. Cell Rep. 7, 1077–1092 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Chen, S. X., Kim, A. N., Peters, A. J. & Komiyama, T. Subtype-specific plasticity of inhibitory circuits in motor cortex during motor learning. Nat. Neurosci. 18, 1109–1115 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Padmashri, R., Reiner, B. C., Suresh, A., Spartz, E. & Dunaevsky, A. Altered structural and functional synaptic plasticity with motor skill learning in a mouse model of fragile X syndrome. J. Neurosci. 33, 19715–19723 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Peters, A. J., Chen, S. X. & Komiyama, T. Emergence of reproducible spatiotemporal activity during motor learning. Nature 510, 263–267 (2014).

    Article  CAS  PubMed  Google Scholar 

  13. Peters, A. J., Lee, J., Hedrick, N. G., O’Neil, K. & Komiyama, T. Reorganization of corticospinal output during motor learning. Nat. Neurosci. 20, 1133–1141 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Rioult-Pedotti, M. S., Friedman, D. & Donoghue, J. P. Learning-induced LTP in neocortex. Science 290, 533–536 (2000).

    Article  CAS  PubMed  Google Scholar 

  15. 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 

  16. Yang, G., Pan, F. & Gan, W. B. Stably maintained dendritic spines are associated with lifelong memories. Nature 462, 920–924 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Guo, L. et al. Dynamic rewiring of neural circuits in the motor cortex in mouse models of Parkinson’s disease. Nat. Neurosci. 18, 1299–1309 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Schiemann, J. et al. Cellular mechanisms underlying behavioral state-dependent bidirectional modulation of motor cortex output. Cell Rep. 11, 1319–1330 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Labarrera, C. et al. Adrenergic modulation regulates the dendritic excitability of layer 5 pyramidal neurons in vivo. Cell Rep. 23, 1034–1044 (2018).

    Article  CAS  PubMed  Google Scholar 

  20. Reiner, B. C. & Dunaevsky, A. Deficit in motor training-induced clustering, but not stabilization, of new dendritic spines in FMR1 knock-out mice. PloS ONE 10, e0126572 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  21. Forrest, M. P., Parnell, E. & Penzes, P. Dendritic structural plasticity and neuropsychiatric disease. Nat. Rev. Neurosci. 19, 215–234 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Walsh, J. J. et al. 5-HT release in nucleus accumbens rescues social deficits in mouse autism model. Nature 560, 589–594 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Del Pino, I., Rico, B. & Marin, O. Neural circuit dysfunction in mouse models of neurodevelopmental disorders. Curr. Opin. Neurobiol. 48, 174–182 (2018).

    Article  PubMed  CAS  Google Scholar 

  24. Robinson, J. E. & Gradinaru, V. Dopaminergic dysfunction in neurodevelopmental disorders: recent advances and synergistic technologies to aid basic research. Curr. Opin. Neurobiol. 48, 17–29 (2018).

    Article  CAS  PubMed  Google Scholar 

  25. Goncalves, J. T., Anstey, J. E., Golshani, P. & Portera-Cailliau, C. Circuit level defects in the developing neocortex of fragile X mice. Nat. Neurosci. 16, 903–909 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Nath, T. et al. Using DeepLabCut for 3D markerless pose estimation across species and behaviors. Nat. Protoc. 14, 2152–2176 (2019).

    Article  CAS  PubMed  Google Scholar 

  27. Herzfeld, D. J. & Shadmehr, R. Motor variability is not noise, but grist for the learning mill. Nat. Neurosci. 17, 149–150 (2014).

    Article  CAS  PubMed  Google Scholar 

  28. Lu, H. et al. Loss and gain of MeCP2 cause similar hippocampal circuit dysfunction that is rescued by deep brain stimulation in a Rett syndrome mouse model. Neuron 91, 739–747 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Vitrac, C. & Benoit-Marand, M. Monoaminergic modulation of motor cortex function. Front. Neural Circuits 11, 72 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Sara, S. J. The locus coeruleus and noradrenergic modulation of cognition. Nat. Rev. Neurosci. 10, 211–223 (2009).

    Article  CAS  PubMed  Google Scholar 

  31. Lee, S. H. & Dan, Y. Neuromodulation of brain states. Neuron 76, 209–222 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Conner, J. M., Culberson, A., Packowski, C., Chiba, A. A. & Tuszynski, M. H. Lesions of the basal forebrain cholinergic system impair task acquisition and abolish cortical plasticity associated with motor skill learning. Neuron 38, 819–829 (2003).

    Article  CAS  PubMed  Google Scholar 

  33. Bast, N., Poustka, L. & Freitag, C. M. The locus coeruleus-norepinephrine system as pacemaker of attention—a developmental mechanism of derailed attentional function in autism spectrum disorder. Eur. J. Neurosci. 47, 115–125 (2018).

    Article  PubMed  Google Scholar 

  34. Armbruster, B. N., Li, X., Pausch, M. H., Herlitze, S. & Roth, B. L. Evolving the lock to fit the key to create a family of G protein-coupled receptors potently activated by an inert ligand. Proc. Natl Acad. Sci. USA 104, 5163–5168 (2007).

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  35. Sciolino, N. R. et al. Recombinase-dependent mouse lines for chemogenetic activation of genetically defined cell types. Cell Rep. 15, 2563–2573 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Fortress, A. M. et al. Designer receptors enhance memory in a mouse model of Down syndrome. J. Neurosci. 35, 1343–1353 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Martins, A. R. & Froemke, R. C. Coordinated forms of noradrenergic plasticity in the locus coeruleus and primary auditory cortex. Nat. Neurosci. 18, 1483–1492 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Breton-Provencher, V. & Sur, M. Active control of arousal by a locus coeruleus GABAergic circuit. Nat. Neurosci. 22, 218–228 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Aston-Jones, G. & Waterhouse, B. Locus coeruleus: from global projection system to adaptive regulation of behavior. Brain Res. 1645, 75–78 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Zhu, Y., Wienecke, C. F., Nachtrab, G. & Chen, X. A thalamic input to the nucleus accumbens mediates opiate dependence. Nature 530, 219–222 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Memmesheimer, R. M., Rubin, R., Olveczky, B. P. & Sompolinsky, H. Learning precisely timed spikes. Neuron 82, 925–938 (2014).

    Article  CAS  PubMed  Google Scholar 

  42. Olshausen, B. A. & Field, D. J. Sparse coding of sensory inputs. Curr. Opin. Neurobiol. 14, 481–487 (2004).

    Article  CAS  PubMed  Google Scholar 

  43. Palm, G. Neural associative memories and sparse coding. Neural Netw. 37, 165–171 (2013).

    Article  PubMed  Google Scholar 

  44. McIntyre, C. K., McGaugh, J. L. & Williams, C. L. Interacting brain systems modulate memory consolidation. Neurosci. Biobehav. Rev. 36, 1750–1762 (2012).

    Article  PubMed  Google Scholar 

  45. Arnsten, A. F., Wang, M. J. & Paspalas, C. D. Neuromodulation of thought: flexibilities and vulnerabilities in prefrontal cortical network synapses. Neuron 76, 223–239 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. McCormick, D. A., Pape, H. C. & Williamson, A. Actions of norepinephrine in the cerebral cortex and thalamus: implications for function of the central noradrenergic system. Prog. Brain Res. 88, 293–305 (1991).

    Article  CAS  PubMed  Google Scholar 

  47. Stein, I. S. & Zito, K. Dendritic spine elimination: molecular mechanisms and implications. Neuroscientist 25, 27–47 (2019).

    Article  CAS  PubMed  Google Scholar 

  48. Piochon, C. et al. Cerebellar plasticity and motor learning deficits in a copy-number variation mouse model of autism. Nat. Commun. 5, 5586 (2014).

    Article  CAS  PubMed  Google Scholar 

  49. Uematsu, A. et al. Modular organization of the brainstem noradrenaline system coordinates opposing learning states. Nat. Neurosci. 20, 1602–1611 (2017).

    Article  CAS  PubMed  Google Scholar 

  50. Beas, B. S. et al. The locus coeruleus drives disinhibition in the midline thalamus via a dopaminergic mechanism. Nat. Neurosci. 21, 963–973 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Stubbusch, J. et al. Generation of the tamoxifen-inducible DBH-Cre transgenic mouse line DBH-CT. Genesis 49, 935–941 (2011).

    Article  CAS  PubMed  Google Scholar 

  52. Ayling, O. G., Harrison, T. C., Boyd, J. D., Goroshkov, A. & Murphy, T. H. Automated light-based mapping of motor cortex by photoactivation of channelrhodopsin-2 transgenic mice. Nat. Methods 6, 219–224 (2009).

    Article  CAS  PubMed  Google Scholar 

  53. Chen, S. X. et al. The transcription factor MEF2 directs developmental visually driven functional and structural metaplasticity. Cell 151, 41–55 (2012).

    Article  CAS  PubMed  Google Scholar 

  54. Holtmaat, A. et al. Long-term, high-resolution imaging in the mouse neocortex through a chronic cranial window. Nat. Protoc. 4, 1128–1144 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Suresh, A. & Dunaevsky, A. Relationship between synaptic AMPAR and spine dynamics: impairments in the FXS mouse. Cereb. Cortex 27, 4244–4256 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank B. Lacoste for sharing the 16p11.2+/ mice; B. Liu, J.C. Beique, M. Rousseaux, R. Naud and members of the Chen laboratory for comments and discussions; and B.H. Liu, H. Kato, T. Bui and M. Rousseaux for providing feedback on the manuscript. This work was supported by grants for S.X.C. from the Canada Research Chair (CRC) (grant no. 950-231274), the Canadian Institutes of Health Research (CIHR) (grant no. 153254), the Natural Sciences and Engineering Research Council of Canada (NSERC) (grant no. 05308), the Simons Foundation, the Scottish Rite Charitable Foundation, and the Brain & Behavior Research Foundation (NARSAD) (grant no. 27177).

Author information

Authors and Affiliations

Authors

Contributions

X.Y. and S.X.C. conceived the project. All the experiments were performed by X.Y. with help from N.J., J.Y., N.A., M.-E.M. and L.C. N.J. collected the behavior data and performed the DeepLabCut analyses. J.W. collected and analyzed the data for the colocalization experiments of LC-NA neurons. N.A. collected and analyzed the data for ChAT-positive axons in motor cortex, and collected and analyzed the behavior data for the administration of muscimol and noradrenergic receptor antagonists. M.-E.M. collected and analyzed the data for NET-positive axons in motor cortex. L.C. helped with collecting the calcium imaging data and calcium imaging data analysis. All the rest of the data were collected and analyzed by X.Y. under supervision of S.X.C. X.Y., N.J. and S.X.C. wrote the manuscript.

Corresponding author

Correspondence to Simon X. Chen.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Neuroscience thanks the anonymous reviewers for their contribution to the peer review of this work.

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

Extended data

Extended Data Fig. 1 Mice showed a decrease in movement variation with motor learning in the rotating-disk task.

a, Schematic of the rotating-disk task recorded by a high speed and resolution camera from the top. b-c, DeepLabCut was used to unbiasedly track the position of head (white), neck (black), thorax (white), and sacrum (orange) across 12 sessions. Representative plots of all the corporeal coordinates from the neck and sacrum and the calculated angles from one RE of a mouse in session 1 (b) and session 12 (c). a.u., arbitrary unit. d, Mean relative variance of the angles decreased with motor learning (F(3,20) = 5.96, P = 0.004; 1-way ANOVA; Solid line, mean variance; shading, s.e.m.; n = 5 mice). e, Mean relative variance of the angles in ‘early’ (first 1/3 of the session) and ‘late’ (last 1/3 of the session) groups in session 1, 4, 8, and 12. Both ‘early’ and ‘late’ groups were comparable to each other throughout learning (n = 5 mice). One-tailed bootstrap test. Error bars indicate s.e.m.

Extended Data Fig. 2 WT and 16p11.2+/− mice showed similar performance on the head-fixed rotating-disk task in session 1.

a, Schematic of the rotating-disk task. b-c, Mean running attempts (b), distance travelled (c) in session 1. Both groups showed a comparable number of running attempts but had low total distance travelled (WT, n = 10 mice; 16p11.2+/, n = 7 mice). One-tailed bootstrap test. d-e, Mean RE counts (d) and RE duration (e) in session 1. Both groups showed comparable performance in session 1 (WT, n = 10 mice; 16p11.2+/, n = 7 mice). One-tailed bootstrap test. f, Mean probability distribution of sub-threshold velocities during SEs in session 1 (n = 7 mice per group). Both groups showed a comparable distribution of velocities. Error bars indicate s.e.m.

Extended Data Fig. 3 The elevated ensemble activity is RE-specific, which only occurs during the initial learning phase in 16p11.2+/− mice and is not related with the number of RE-active neurons or the ensemble activity during SEs.

a, Schematic of the two-photon Ca2+ imaging on the rotating-disk task. b, Representative of all the regions of interests (ROIs) drawn from one mouse in session 12. Each ROI represents a neuron. The mean ROIs analyzed in session 12 was 400 ± 48 neurons. Scale bar = 100 µm. c, Frequency of synchrony events from ‘pre-existing RE-active neurons’ in session 1-3 (same population as Fig. 2e) and followed throughout learning. ‘Pre-existing RE-active neurons’ in 16p11.2+/ mice initially showed more synchrony events, which desynchronized with training (F(3,20) = 3.27, P= 0.043, 1-way ANOVA). In contrast, ‘pre-existing RE-active neurons’ in WT mice had low synchrony and remained constant throughout learning (F(3,20) = 0.14, P= 0.933, 1-way ANOVA). One-tailed bootstrap test. d, Median activity index from ‘all RE-active neurons’ in session 1-3, 4-6, 7-9, and 10-12. Only ‘all RE-active neurons’ in 16p11.2+/ mice initially (session 1–3) showed higher activity than WT mice. One-tailed bootstrap test. e, Frequency of synchrony events from ‘all RE-active neurons’ in session 1–3, 4–6, 7–9, and 10–12 (same population as d). Only ‘all RE-active neurons’ in 16p11.2+/ mice initially (session 1–3) showed elevated synchrony events compared to WT mice. One-tailed bootstrap test. f, The number of ‘all RE-active neurons’ across training was similar between WT and 16p11.2+/ mice. One-tailed bootstrap test. g, The number of ‘newly-emerged RE-active neurons’ across training was similar between WT and 16p11.2+/ mice. One-tailed bootstrap test. h, Mean ΔF/F0 during SEs from all RE-active neurons in session 1–3. One-tailed bootstrap test. i, Median activity index of RE-active neurons in session 1–3, calculated between sub-threshold events and stationary events during SEs. One-tailed bootstrap test. j, k, Median activity index (j) and synchrony events (k) from SE-active neurons in session 1–3. Both WT and 16p11.2+/ mice showed a similar level of activity and synchrony events. One-tailed bootstrap test. l, The number of ‘newly-emerged RE-active neurons’ during training (session 4–6) was not correlated with the improvement in task performance (session 4–7) (Pearson linear correlation coefficient, r = −0.304, P = 0.337). m, Schematic of the two-photon spine imaging on the rotating-disk task. n, Low magnification images for the L2/3 distal dendrites (left) and somatic (right) regions of L2/3 excitatory neurons. Scale bar = 50 µm. **P < 0.01, ***P < 0.001 Error bars indicate s.e.m.

Extended Data Fig. 4 16p11.2+/− mice retain the memory of the head-fixed rotating-disk task.

a, Left: schematic of the head-fixed rotating-disk task. Right: Experimental timeline. Orange lines indicate the spine imaging sessions during retraining. P, postnatal day. b–d, Behavioral performances in the last session of training (Training_S12), the first session of retraining (Retraining_S1), and the last session of retraining (Retraining_S7). Both groups retained the motor memory and had comparable mean total running distance travelled (b), RE counts (c), and RE duration (d) between all 3 sessions (WT, n = 6 mice; 16p11.2+/, n = 5 mice). One-tailed bootstrap test with Bonferroni correction. e-g, Mean spine density (e), pre-existing spine elimination (f), and spine dynamics between every 3 sessions during retraining (g) showed no significant changes (WT, n = 3 mice, 136 spines; 16p11.2+/, n = 3 mice, 154 spines). 2-way ANOVA. Error bars indicate s.e.m.

Extended Data Fig. 5 Number of VTA-DA and BF-ACh activated neurons are not affected in 16p11.2+/− mice.

a, b, Representative coronal sections with immunostaining from 1-day rotarod trained mice. a, Left: a coronal section stained for c-Fos (green) and tyrosine hydroxylase (TH, red) (~3.3 mm posterior from the bregma). VTA and hippocampus are emphasized by dotted boxes. Scale bar = 1,000 µm. Middle: example images of c-Fos+ cells, TH+ cells, and co-localized cells in the VTA and hippocampus in WT and 16p11.2+/ mice. Scale bar = 100 µm. Right: mean percentage of c-Fos+ cells among the dopaminergic neurons in VTA. A comparable level of c-Fos+-TH+ neurons was found between the two groups (n = 6 mice). One-tailed bootstrap test. b, Left: a coronal section stained for c-Fos (green) and choline acetyltransferase (ChAT, red) (~0.9 mm posterior from the bregma). Basal forebrain and hypothalamus are emphasized by dotted boxes; scale bar = 1,000 µm. Middle: example images of c-Fos+ cells, ChAT+ cells, and co-localized cells in the basal forebrain and hypothalamus in WT and 16p11.2+/ mice. Scale bar = 100 µm. Right: mean percentage of c-Fos+ cells among the cholinergic neurons in basal forebrain. A comparable level of c-Fos+-ChAT+ neurons was found between the two groups (n = 6 mice). One-tailed bootstrap test. c, Mean DAPI density in the L2/3 of the motor cortex. No significant difference was found between the two groups (n = 5 mice). One-tailed bootstrap test. d, ChAT staining to measure cholinergic axonal innervation in the L2/3 of the motor cortex. Mean innervation density showed there was no significant difference between the two groups (n = 3 mice). One-tailed bootstrap test. e, NET staining to measure noradrenergic axonal innervation in the L2/3 of the somatosensory cortex. Mean innervation density showed there was no significant difference between the two groups (n = 6 mice). One-tailed bootstrap test. Error bars indicate s.e.m.

Extended Data Fig. 6 Pharmacogenetic activation restored the number of activated LC-NA neurons in 16p11.2+/− mice.

a, Left: schematic of the brain section used to stain for activated c-Fos and NA neurons in the LC. LC is emphasized by the orange shadow. Right: example images of hM3Dq-mCherry (far left), TH+ (center left), c-Fos+ (center right), and co-localized (far right) in the LC-NA neurons from 16p11.2+/::DBH-CreERT+/ mice expressing hM3Dq-mCherry. Scale bar = 50 µm. Displayed images are representative of experiments repeated in five mice with similar results obtained. b, Mean percentage of LC-NA neurons expressing hM3Dq-mCherry among the total LC-NA neurons. c, Mean percentage of NA neurons that are c-Fos+ in WT mice (n = 6 mice), 16p11.2+/ mice (n = 5 mice), and hM3Dq group (n = 5 mice) after one session of the rotating-disk training. One-tailed bootstrap test with Bonferroni correction. d, Mean total number of LC-NA neurons in WT mice (n = 6 mice) and 16p11.2+/ mice (n = 6 mice). One-tailed bootstrap test. e, Schematic of the head-fixed rotating-disk task. f–h, Mean total distance travelled (f), RE counts (g), and RE duration (h) in session 1 of the four experimental groups in the DREADDs experiments. All four groups showed comparable performance. One-tailed bootstrap test. **P < 0.01 Error bars indicate s.e.m.

Extended Data Fig. 7 Proposed model for circuit dysfunctions during delayed motor learning in 16p11.2+/− mice.

Left: motor learning induces circuit reorganizations in the motor cortex through a multi-step process: In WT mice, during the initial phase of learning, RE neurons explore learning-related activity patterns, and nascent spines are formed to create structural substrates for the new motor memory (1). As learning progresses, NA is being released and ‘task-related’ RE-active neurons refine learning-related activity patterns (2). The refinement process to generate learned movement leads to the elimination of unnecessary spines to achieve proper circuit reorganization (3). Right: in the 16p11.2+/ mice, the process of circuit reorganization is delayed. The low levels of NA in the motor cortex of 16p11.2+/ mice could alter the baseline excitability of neurons. Hence, during an RE in the initial phase of learning, ‘task-related’ RE-active neurons become highly active, which leads to elevated ensemble synchrony but learning-induced spine formation is not affected (1). As learning progresses, NA is being released, and neuronal excitability reverts to baseline levels. As the ensemble begins to desynchronize (2), it allows ‘task-related’ RE-active neurons to refine learning-related activity patterns (3). However, the extra step of lowering neuronal activity and desynchronization delays the circuit’s ability to refine activity patterns and eliminate unnecessary spines to achieve proper circuit reorganization (4); thereby causing the delayed motor learning.

Supplementary information

Supplementary Information

Supplementary Table 1. Summary of the number of mice, spines, branch number, branch length and density in all experimental conditions.

Reporting Summary

Supplementary Video 1

WT mouse underwent head-fixed rotating-disk training in session 1. The mouse showed multiple movement attempts but was not able to smoothly control the rotation of the bi-directional disk under head-fixation.

Supplementary Video 2

WT mouse underwent head-fixed rotating-disk training in session 12. The mouse learned to modify its gait strategy to reduce movement variability and facilitate continuous running on the disk.

Supplementary Video 3

In vivo two-photon imaging of LC-NA axons expressing hM3Dq-mCherry in the motor cortex 4 weeks after the viral injection. The stack of images was taken at 512 × 512 pixels covering 503 × 429 µm2 with a 1-µm z-axis step size (100–208 µm from the pial surface) between planes. The mCherry-expressing axons were abundant and covered almost the entire imaging plane.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yin, X., Jones, N., Yang, J. et al. Delayed motor learning in a 16p11.2 deletion mouse model of autism is rescued by locus coeruleus activation. Nat Neurosci 24, 646–657 (2021). https://doi.org/10.1038/s41593-021-00815-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-021-00815-7

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