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:

Early hyperactivity and precocious maturation of corticostriatal circuits in Shank3B−/− mice

Abstract

Some autistic individuals exhibit abnormal development of the caudate nucleus and associative cortical areas, suggesting potential dysfunction of cortico-basal ganglia (BG) circuits. Using optogenetic and electrophysiological approaches in mice, we identified a narrow postnatal period that is characterized by extensive glutamatergic synaptogenesis in striatal spiny projection neurons (SPNs) and a concomitant increase in corticostriatal circuit activity. SPNs during early development have high intrinsic excitability and respond strongly to cortical afferents despite sparse excitatory inputs. As a result, striatum and corticostriatal connectivity are highly sensitive to acute and chronic changes in cortical activity, suggesting that early imbalances in cortical function alter BG development. Indeed, a mouse model of autism with deletions in Shank3 (Shank3B−/−) shows early cortical hyperactivity, which triggers increased SPN excitatory synapse and corticostriatal hyperconnectivity. These results indicate that there is a tight functional coupling between cortex and striatum during early postnatal development and suggest a potential common circuit dysfunction that is caused by cortical hyperactivity.

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

Figure 1: Rapid development of striatal SPN excitatory input in mice after P10.
Figure 2: Correlated increase in cortical and striatal activity in vivo from P10 to P16.
Figure 3: Corticostriatal coupling during early development.
Figure 4: Hyperexcitability of SPNs during early development.
Figure 5: Precocious maturation of striatal glutamatergic inputs in Shank3B−/− SPNs.
Figure 6: Cortical hyperactivity in neonatal Shank3B−/− mice.
Figure 7: Elevated cortical activity during early development increases corticostriatal connectivity.
Figure 8: Early increase in corticostriatal drive in Shank3B−/− mice is a result of cortical hyperactivity.

Similar content being viewed by others

References

  1. Shepherd, G.M.G. Corticostriatal connectivity and its role in disease. Nat. Rev. Neurosci. 14, 278–291 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Oldenburg, I.A. & Sabatini, B.L. Antagonistic but not symmetric regulation of primary motor cortex by basal ganglia direct and indirect pathways. Neuron 86, 1174–1181 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Difiglia, M., Pasik, P. & Pasik, T. Early postnatal development of the monkey neostriatum: a Golgi and ultrastructural study. J. Comp. Neurol. 190, 303–331 (1980).

    Article  CAS  PubMed  Google Scholar 

  4. Levine, M.S., Fisher, R.S., Hull, C.D. & Buchwald, N.A. Postnatal development of identified medium-sized caudate spiny neurons in the cat. Brain Res. 389, 47–62 (1986).

    Article  CAS  PubMed  Google Scholar 

  5. Tepper, J.M., Sharpe, N.A., Koós, T.Z. & Trent, F. Postnatal development of the rat neostriatum: electrophysiological, light- and electron-microscopic studies. Dev. Neurosci. 20, 125–145 (1998).

    Article  CAS  PubMed  Google Scholar 

  6. Kozorovitskiy, Y., Saunders, A., Johnson, C.a., Lowell, B.B. & Sabatini, B.L. Corrigendum: Recurrent network activity drives striatal synaptogenesis. Nature 489, 326–326 (2012).

    Article  CAS  Google Scholar 

  7. Langen, M., Durston, S., Staal, W.G., Palmen, S.J.M.C. & van Engeland, H. Caudate nucleus is enlarged in high-functioning medication-naive subjects with autism. Biol. Psychiatry 62, 262–266 (2007).

    Article  PubMed  Google Scholar 

  8. Wolff, J.J., Hazlett, H.C., Lightbody, A.A., Reiss, A.L. & Piven, J. Repetitive and self-injurious behaviors: associations with caudate volume in autism and fragile X syndrome. J. Neurodev. Disord. 5, 12 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Langen, M. et al. Changes in the developmental trajectories of striatum in autism. Biol. Psychiatry 66, 327–333 (2009).

    Article  PubMed  Google Scholar 

  10. Langen, M. et al. Changes in the development of striatum are involved in repetitive behavior in autism. Biol. Psychiatry 76, 405–411 (2014).

    Article  PubMed  Google Scholar 

  11. Lord, C., Cook, E.H., Leventhal, B.L. & Amaral, D.G. Autism spectrum disorders. Neuron 28, 355–363 (2000).

    Article  CAS  PubMed  Google Scholar 

  12. DeLong, M. & Wichmann, T. Changing views of basal ganglia circuits and circuit disorders. Clin. EEG Neurosci. 41, 61–67 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Abrahams, B.S. & Geschwind, D.H. Advances in autism genetics: on the threshold of a new neurobiology. Nat. Rev. Genet. 9, 341–355 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Jiang, Y.H. & Ehlers, M.D. Modeling autism by SHANK gene mutations in mice. Neuron 78, 8–27 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Durand, C.M. et al. Mutations in the gene encoding the synaptic scaffolding protein SHANK3 are associated with autism spectrum disorders. Nat. Genet. 39, 25–27 (2007).

    Article  CAS  PubMed  Google Scholar 

  16. Gauthier, J. et al. Novel de novo SHANK3 mutation in autistic patients. Am. J. Med. Genet. B. Neuropsychiatr. Genet. 150B, 421–424 (2009).

    Article  CAS  PubMed  Google Scholar 

  17. Roussignol, G. et al. Shank expression is sufficient to induce functional dendritic spine synapses in aspiny neurons. J. Neurosci. 25, 3560–3570 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Arons, M.H. et al. Autism-associated mutations in ProSAP2/Shank3 impair synaptic transmission and neurexin-neuroligin-mediated transsynaptic signaling. J. Neurosci. 32, 14966–14978 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Han, K. et al. SHANK3 overexpression causes manic-like behavior with unique pharmacogenetic properties. Nature 503, 72–77 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Verpelli, C. et al. Importance of Shank3 protein in regulating metabotropic glutamate receptor 5 (mGluR5) expression and signaling at synapses. J. Biol. Chem. 286, 34839–34850 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Bozdagi, O. et al. Haploinsufficiency of the autism-associated Shank3 gene leads to deficits in synaptic function, social interaction and social communication. Mol. Autism 1, 15 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Yang, M. et al. Reduced excitatory neurotransmission and mild autism-relevant phenotypes in adolescent Shank3 null mutant mice. J. Neurosci. 32, 6525–6541 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Peça, J. et al. Shank3 mutant mice display autistic-like behaviours and striatal dysfunction. Nature 472, 437–442 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Wang, X. et al. Synaptic dysfunction and abnormal behaviors in mice lacking major isoforms of Shank3. Hum. Mol. Genet. 20, 3093–3108 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Harris, J.A. et al. Anatomical characterization of Cre driver mice for neural circuit mapping and manipulation. Front. Neural Circuits 8, 76 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Petralia, R.S. et al. Selective acquisition of AMPA receptors over postnatal development suggests a molecular basis for silent synapses. Nat. Neurosci. 2, 31–36 (1999).

    Article  CAS  PubMed  Google Scholar 

  27. Busetto, G., Higley, M.J. & Sabatini, B.L. Developmental presence and disappearance of postsynaptically silent synapses on dendritic spines of rat layer 2/3 pyramidal neurons. J. Physiol. (Lond.) 586, 1519–1527 (2008).

    Article  CAS  Google Scholar 

  28. Khazipov, R. et al. Early motor activity drives spindle bursts in the developing somatosensory cortex. Nature 432, 758–761 (2004).

    Article  CAS  PubMed  Google Scholar 

  29. Carter, A.G., Soler-Llavina, G.J. & Sabatini, B.L. Timing and location of synaptic inputs determine modes of subthreshold integration in striatal medium spiny neurons. J. Neurosci. 27, 8967–8977 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Choi, S. & Lovinger, D.M. Decreased probability of neurotransmitter release underlies striatal long-term depression and postnatal development of corticostriatal synapses. Proc. Natl. Acad. Sci. USA 94, 2665–2670 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Moody, W.J. & Bosma, M.M. Ion channel development, spontaneous activity, and activity-dependent development in nerve and muscle cells. Physiol. Rev. 85, 883–941 (2005).

    Article  CAS  PubMed  Google Scholar 

  32. Kozorovitskiy, Y. et al. Neuromodulation of excitatory synaptogenesis in striatal development. eLife 4, e10111 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  33. Sohur, U.S., Padmanabhan, H.K., Kotchetkov, I.S., Menezes, J.R.L. & Macklis, J.D. Anatomic and molecular development of corticostriatal projection neurons in mice. Cereb. Cortex 24, 293–303 (2014).

    Article  PubMed  Google Scholar 

  34. Colonnese, M.T. et al. A conserved switch in sensory processing prepares developing neocortex for vision. Neuron 67, 480–498 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Kilb, W., Kirischuk, S. & Luhmann, H.J. Electrical activity patterns and the functional maturation of the neocortex. Eur. J. Neurosci. 34, 1677–1686 (2011).

    Article  PubMed  Google Scholar 

  36. Etherington, S.J. & Williams, S.R. Postnatal development of intrinsic and synaptic properties transforms signaling in the layer 5 excitatory neural network of the visual cortex. J. Neurosci. 31, 9526–9537 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Gertler, T.S., Chan, C.S. & Surmeier, D.J. Dichotomous anatomical properties of adult striatal medium spiny neurons. J. Neurosci. 28, 10814–10824 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Kirkby, L.A., Sack, G.S., Firl, A. & Feller, M.B. A role for correlated spontaneous activity in the assembly of neural circuits. Neuron 80, 1129–1144 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Grabrucker, A.M. et al. Concerted action of zinc and ProSAP/Shank in synaptogenesis and synapse maturation. EMBO J. 30, 569–581 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Gogolla, N., Takesian, A.E., Feng, G., Fagiolini, M. & Hensch, T.K. Sensory integration in mouse insular cortex reflects GABA circuit maturation. Neuron 83, 894–905 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Martin, L.J. & Cork, L.C. The non-human primate striatum undergoes marked prolonged remodeling during postnatal development. Front. Cell. Neurosci. 8, 294 (2014).

    PubMed  PubMed Central  Google Scholar 

  42. Lewine, J.D. et al. Magnetoencephalographic patterns of epileptiform activity in children with regressive autism spectrum disorders. Pediatrics 104, 405–418 (1999).

    Article  CAS  PubMed  Google Scholar 

  43. Gonçalves, 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  PubMed  PubMed Central  CAS  Google Scholar 

  44. Bateup, H.S. et al. Excitatory/inhibitory synaptic imbalance leads to hippocampal hyperexcitability in mouse models of tuberous sclerosis. Neuron 78, 510–522 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Peñagarikano, O. et al. Absence of CNTNAP2 leads to epilepsy, neuronal migration abnormalities, and core autism-related deficits. Cell 147, 235–246 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Han, S. et al. Autistic-like behaviour in Scn1a+/− mice and rescue by enhanced GABA-mediated neurotransmission. Nature 489, 385–390 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Wolff, J.J. et al. Longitudinal patterns of repetitive behavior in toddlers with autism. J. Child Psychol. Psychiatry 55, 945–953 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  48. Goldman, S. et al. Motor stereotypies in children with autism and other developmental disorders. Dev. Med. Child Neurol. 51, 30–38 (2009).

    Article  PubMed  Google Scholar 

  49. Goldman, S. & Greene, P.E. Stereotypies in autism: a video demonstration of their clinical variability. Front. Integr. Neurosci. 6, 121 (2012).

    PubMed  Google Scholar 

  50. Harris, K.M., Mahone, E.M. & Singer, H.S. Nonautistic motor stereotypies: clinical features and longitudinal follow-up. Pediatr. Neurol. 38, 267–272 (2008).

    Article  PubMed  Google Scholar 

  51. Legéndy, C.R. & Salcman, M. Bursts and recurrences of bursts in the spike trains of spontaneously active striate cortex neurons. J. Neurophysiol. 53, 926–939 (1985).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We thank I. Oldenburg for help with in vivo recordings and analysis and J. Levasseur and R. Pemberton for mouse genotyping and colony management. We thank S. da Silva, C. Deister and the members of the Sabatini laboratory for helpful discussions and critical reading of the manuscript. R.T.P. was supported by the Alice and Joseph Brooks fellowship and the Nancy Lurie Marks clinical and research fellowship in autism. Y.K. was supported by the Leonard and Isabelle Goldenson Research Fellowship and the Nancy Lurie Marks Family Foundation. This work was supported by the National Institute of Neurological Disorders and Stroke (NS046579, to B.L.S.) and the Nancy Lurie Marks Foundation.

Author information

Authors and Affiliations

Authors

Contributions

R.T.P. and B.L.S. conceived the study and wrote the manuscript. R.T.P. carried out in vivo recordings and analyzed the data. R.T.P., W.W. and Y.K. carried out in vitro slice recordings and R.T.P. analyzed the data. D.M.C. performed the behavioral experiments and dendritic spine imaging and analysis.

Corresponding author

Correspondence to Bernardo L Sabatini.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Correlated oEPSC peak amplitude recorded at Vh=−70 and −20 mV.

Average oEPSC peak amplitude recorded in P14 SPNs of Rbp4-Cre;ChR2f/f mice at Vh=−70 and −20 mV. Note the linear relationship between oEPSCs recorded under the two conditions across a wide range of oEPSC amplitudes.

Supplementary Figure 2 Average cortical and striatal multi-unit responses to optogenetic stimulation of Rbp4-Cre;ChR2f/f mice.

(a) Peri-stimulus time histogram (5 ms bin) of APs fired by P10−11 (black) and P14−16 (red) cortical and (b) striatal units in response to extracranial optogenetic stimulation of Rbp4-Cre;ChR2f/f mice with 100 pulses of 473 nm light (blue rectangle). Shaded regions represent ± SEM. Arrows point to time bin with peak firing rate. Note the 5 ms shift in response latency of striatal units from P10−11 to P14−16.

Supplementary Figure 3 Similar presynaptic release properties in WT and Shank3B−/ SPNs across development.

(a) Example traces of eEPSCs evoked in SPNs of P14 WT and Shank3B−/ mice in response to paired electrical pulses (P1 and P2) with 50 ms ISI. (b) Mean ± SEM ratio of eEPSC amplitude in response to the paired stimulation pulses (P2/P1) in P14 and (c) P9 WT and Shank3B−/ SPNs.

Supplementary Figure 4 Increased Rbp4+ excitatory input onto Shank3B−/− SPNs during early development.

(a) Experimental diagram depicting whole cell recordings in P14 SPNs of dorsomedial striatum in acute brain slices of Shank3B−/;Rbp4-Cre;ChR2-YFPf/wt mice and optogenetic fiber stimulation using whole field illumination (blue cone). Scale bar, 1 mm. (b) Representative traces of AMPAR oEPSCs recorded in SPNs of Shank3B+/ or Shank3B−/ mice under voltage clamp (Vh= -70 mV) in response to brief pulses of 473 nm laser light (blue rectangle). (c) Mean oEPSC peak amplitude ± SEM recorded in Shank3B+/ or Shank3B−/ SPNs. (d) Pair-wise comparison of average oEPSC amplitude in animals recorded in (c). Note that SPNs from Shank3B−/ animals have consistently larger oEPSC amplitude compared to SPNs from Shank3B+/ heterozygous littermates.

Supplementary Figure 5 Similar intrinsic excitability of WT and Shank3B−/ SPNs at P13−14.

(a) Mean ± SEM current-voltage (I-V) relationship in WT and KO SPNs. (b) Mean spike threshold potential (c) rheobase current (d) Vrest and (e) Current-firing rate (I-F) plot of WT and KO SPNs recorded at P13−14. Error bars represent ± SEM.

Supplementary Figure 6 Decreased locomotion of AAV-Cre injected vGATf/f mice.

(a) Heat map representing locomotion of vGATf/f mice (control) and vGATf/f littermates injected with AAV-Cre-EGFP (Cre) in an open chamber for 15 min. Color scale represents normalized time spent at each location. Scale bar, 10 cm. (b) Mean average velocity ± SEM of control and Cre injected vGATf/f animals. (c) Mean average time moving ± SEM of control and Cre injected vGATf/f animals. (d) Mean average total distance moved ± SEM of control and Cre injected vGATf/f animals.

Supplementary Figure 7 Similar PPR of eEPSC in SPNs of vGATf/f (control) and vGATf/f mice injected with AAV-Cre-EGFP.

(a) Example traces of eEPSCs evoked in SPNs of control and Cre-injected littermates in response to paired electrical pulses (P1 and P2) with 50 ms ISI. (b) Mean ± SEM ratio of eEPSC amplitude in response to the paired stimulation pulses (P2/P1).

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Peixoto, R., Wang, W., Croney, D. et al. Early hyperactivity and precocious maturation of corticostriatal circuits in Shank3B−/− mice. Nat Neurosci 19, 716–724 (2016). https://doi.org/10.1038/nn.4260

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4260

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