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:

Social deficits in Shank3-deficient mouse models of autism are rescued by histone deacetylase (HDAC) inhibition

A Publisher Correction to this article was published on 05 June 2018

This article has been updated

Abstract

Haploinsufficiency of the SHANK3 gene is causally linked to autism spectrum disorder (ASD), and ASD-associated genes are also enriched for chromatin remodelers. Here we found that brief treatment with romidepsin, a highly potent class I histone deacetylase (HDAC) inhibitor, alleviated social deficits in Shank3-deficient mice, which persisted for ~3 weeks. HDAC2 transcription was upregulated in these mice, and knockdown of HDAC2 in prefrontal cortex also rescued their social deficits. Nuclear localization of β-catenin, a Shank3-binding protein that regulates cell adhesion and transcription, was increased in Shank3-deficient mice, which induced HDAC2 upregulation and social deficits. At the downstream molecular level, romidepsin treatment elevated the expression and histone acetylation of Grin2a and actin-regulatory genes and restored NMDA-receptor function and actin filaments in Shank3-deficient mice. Taken together, these findings highlight an epigenetic mechanism underlying social deficits linked to Shank3 deficiency, which may suggest potential therapeutic strategies for ASD patients bearing SHANK3 mutations.

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: Treatment with the HDAC inhibitor romidepsin induces robust, prolonged rescue of autism-like social deficits in Shank3-deficient mice, while a variety of drugs for psychiatric disorders fail to do so.
Fig. 2: Romidepsin treatment does not affect locomotion, anxiety-like behavior or neuronal survival in Shank3-deficient mice.
Fig. 3: Shank3 deficiency induces HDAC2 upregulation, and HDAC2 knockdown in PFC rescues autism-like social deficits.
Fig. 4: β-catenin is nuclear-translocated to activate HDAC2 transcription in Shank3-deficient mice, and manipulation of β-catenin directly affects HDAC2 transcription and social behaviors.
Fig. 5: Romidepsin treatment increases Grin2a transcription and histone acetylation and restores NMDAR synaptic function in PFC of Shank3-deficient mice.
Fig. 6: Romidepsin treatment elevates the expression of actin regulators and normalizes actin filaments in PFC of Shank3-deficient mice.
Fig. 7: Romidepsin treatment induces genome-wide restoration or elevation of genes involved in neural signaling in PFC of Shank3-deficient mice.

Similar content being viewed by others

Change history

  • 05 June 2018

    In the version of this article initially published, the blue diamonds in Fig. 2a–d were defined as Shank3+/Δc + saline; the correct definition is Shank3+/Δc + RMD. The error has been corrected in the HTML and PDF versions of the article.

References

  1. Silverman, J. L., Yang, M., Lord, C. & Crawley, J. N. Behavioural phenotyping assays for mouse models of autism. Nat. Rev. Neurosci. 11, 490–502 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Naisbitt, S. et al. Shank, a novel family of postsynaptic density proteins that binds to the NMDA receptor/PSD-95/GKAP complex and cortactin. Neuron 23, 569–582 (1999).

    CAS  PubMed  Google Scholar 

  3. Bonaglia, M. C. et al. Disruption of the ProSAP2 gene in a t(12;22)(q24.1;q13.3) is associated with the 22q13.3 deletion syndrome. Am. J. Hum. Genet. 69, 261–268 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. 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).

    CAS  PubMed  Google Scholar 

  5. Betancur, C. & Buxbaum, J. D. SHANK3 haploinsufficiency: a “common” but underdiagnosed highly penetrant monogenic cause of autism spectrum disorders. Mol. Autism 4, 17 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Leblond, C. S. et al. Meta-analysis of SHANK mutations in autism spectrum disorders: a gradient of severity in cognitive impairments. PLoS Genet. 10, e1004580 (2014).

    PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  8. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Kouser, M. et al. Loss of predominant Shank3 isoforms results in hippocampus-dependent impairments in behavior and synaptic transmission. J. Neurosci. 33, 18448–18468 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Duffney, L. J. et al. Autism-like deficits in Shank3-deficient mice are rescued by targeting actin regulators. Cell Rep. 11, 1400–1413 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Wang, X. et al. Altered mGluR5-Homer scaffolds and corticostriatal connectivity in a Shank3 complete knockout model of autism. Nat. Commun. 7, 11459 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Bidinosti, M. et al. CLK2 inhibition ameliorates autistic features associated with SHANK3 deficiency. Science 351, 1199–1203 (2016).

    CAS  PubMed  Google Scholar 

  15. Speed, H. E. et al. Autism-associated insertion mutation (InsG) of SHANK3 exon 21 causes impaired synaptic transmission and behavioral deficits. J. Neurosci. 35, 9648–9665 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Jaramillo, T. C. et al. Altered striatal synaptic function and abnormal behaviour in Shank3 Exon4-9 deletion mouse model of autism. Autism Res. 9, 350–375 (2016).

    PubMed  Google Scholar 

  17. Amodio, D. M. & Frith, C. D. Meeting of minds: the medial frontal cortex and social cognition. Nat. Rev. Neurosci. 7, 268–277 (2006).

    CAS  PubMed  Google Scholar 

  18. Delorme, R. et al. Progress toward treatments for synaptic defects in autism. Nat. Med. 19, 685–694 (2013).

    CAS  PubMed  Google Scholar 

  19. De Rubeis, S. et al. Synaptic, transcriptional and chromatin genes disrupted in autism. Nature 515, 209–215 (2014).

    PubMed  PubMed Central  Google Scholar 

  20. Chen, J. A., Peñagarikano, O., Belgard, T. G., Swarup, V. & Geschwind, D. H. The emerging picture of autism spectrum disorder: genetics and pathology. Annu. Rev. Pathol. 10, 111–144 (2015).

    CAS  PubMed  Google Scholar 

  21. Crawley, J. N., Heyer, W. D. & LaSalle, J. M. Autism and cancer share risk genes, pathways, and drug targets. Trends Genet. 32, 139–146 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Guan, J. S. et al. HDAC2 negatively regulates memory formation and synaptic plasticity. Nature 459, 55–60 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Fischer, A., Sananbenesi, F., Mungenast, A. & Tsai, L. H. Targeting the correct HDAC(s) to treat cognitive disorders. Trends Pharmacol. Sci. 31, 605–617 (2010).

    CAS  PubMed  Google Scholar 

  24. Tsankova, N., Renthal, W., Kumar, A. & Nestler, E. J. Epigenetic regulation in psychiatric disorders. Nat. Rev. Neurosci. 8, 355–367 (2007).

    CAS  PubMed  Google Scholar 

  25. Grayson, D. R., Kundakovic, M. & Sharma, R. P. Is there a future for histone deacetylase inhibitors in the pharmacotherapy of psychiatric disorders? Mol. Pharmacol. 77, 126–135 (2010).

    CAS  PubMed  Google Scholar 

  26. Klimek, V. M. et al. Tolerability, pharmacodynamics, and pharmacokinetics studies of depsipeptide (romidepsin) in patients with acute myelogenous leukemia or advanced myelodysplastic syndromes. Clin. Cancer Res. 14, 826–832 (2008).

    CAS  PubMed  Google Scholar 

  27. Piekarz, R. L. et al. Phase II multi-institutional trial of the histone deacetylase inhibitor romidepsin as monotherapy for patients with cutaneous T-cell lymphoma. J. Clin. Oncol. 27, 5410–5417 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Fraczek, J., Vanhaecke, T. & Rogiers, V. Toxicological and metabolic considerations for histone deacetylase inhibitors. Expert Opin. Drug Metab. Toxicol. 9, 441–457 (2013).

    CAS  PubMed  Google Scholar 

  29. McPheeters, M. L. et al. A systematic review of medical treatments for children with autism spectrum disorders. Pediatrics 127, e1312–e1321 (2011).

    PubMed  Google Scholar 

  30. Valenta, T., Hausmann, G. & Basler, K. The many faces and functions of β-catenin. EMBO J. 31, 2714–2736 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Murase, S., Mosser, E. & Schuman, E. M. Depolarization drives beta-Catenin into neuronal spines promoting changes in synaptic structure and function. Neuron 35, 91–105 (2002).

    CAS  PubMed  Google Scholar 

  32. Yu, X. & Malenka, R. C. Beta-catenin is critical for dendritic morphogenesis. Nat. Neurosci. 6, 1169–1177 (2003).

    CAS  PubMed  Google Scholar 

  33. Stoner, R. et al. Patches of disorganization in the neocortex of children with autism. N. Engl. J. Med. 370, 1209–1219 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Gilman, S. R. et al. Rare de novo variants associated with autism implicate a large functional network of genes involved in formation and function of synapses. Neuron 70, 898–907 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Park, E. et al. The Shank family of postsynaptic density proteins interacts with and promotes synaptic accumulation of the beta PIX guanine nucleotide exchange factor for Rac1 and Cdc42. J. Biol. Chem. 278, 19220–19229 (2003).

    CAS  PubMed  Google Scholar 

  37. Shcheglovitov, A. et al. SHANK3 and IGF1 restore synaptic deficits in neurons from 22q13 deletion syndrome patients. Nature 503, 267–271 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Ebrahimi-Fakhari, D. & Sahin, M. Autism and the synapse: emerging mechanisms and mechanism-based therapies. Curr. Opin. Neurol. 28, 91–102 (2015).

    CAS  PubMed  Google Scholar 

  39. Carlson, G. C. Glutamate receptor dysfunction and drug targets across models of autism spectrum disorders. Pharmacol. Biochem. Behav. 100, 850–854 (2012).

    CAS  PubMed  Google Scholar 

  40. Won, H. et al. Autistic-like social behaviour in Shank2-mutant mice improved by restoring NMDA receptor function. Nature 486, 261–265 (2012).

    CAS  PubMed  Google Scholar 

  41. Chung, W. et al. Social deficits in IRSp53 mutant mice improved by NMDAR and mGluR5 suppression. Nat. Neurosci. 18, 435–443 (2015).

    CAS  PubMed  Google Scholar 

  42. Duffney, L. J. et al. Shank3 deficiency induces NMDA receptor hypofunction via an actin-dependent mechanism. J. Neurosci. 33, 15767–15778 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Yuen, E. Y. et al. Repeated stress causes cognitive impairment by suppressing glutamate receptor expression and function in prefrontal cortex. Neuron 73, 962–977 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Wei, J. et al. Histone modification of Nedd4 ubiquitin ligase controls the loss of AMPA receptors and cognitive impairment induced by repeated stress. J. Neurosci. 36, 2119–2130 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Marcel, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.journal 17, 10–12 (2011).

    Google Scholar 

  46. Kim, D. et al. TopHat2: accurate alignment of transcriptomes in the presence of insertions, deletions and gene fusions. Genome Biol. 14, R36 (2013).

    PubMed  PubMed Central  Google Scholar 

  47. Anders, S., Pyl, P. T. & Huber, W. HTSeq–a Python framework to work with high-throughput sequencing data. Bioinformatics 31, 166–169 (2015).

    CAS  PubMed  Google Scholar 

  48. Robinson, M. D., McCarthy, D. J. & Smyth, G. K. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 26, 139–140 (2010).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank X.Q. Chen for excellent technical support. This work was supported by the Nancy Lurie Marks Family Foundation and grants from the National Institutes of Health (MH112237, MH108842 and DA037618) to Z.Y. We also thank E.F. Trachtman and the Varanasi family for their donations.

Author information

Authors and Affiliations

Authors

Contributions

L.Q. performed immunocytochemical and electrophysiological experiments and analyzed data. K.M. performed behavioral tests and analyzed data. L.Q., Z.-J.W., E.M. and J.W. performed biochemical and molecular biological experiments and analyzed data. Z.H. performed bioinformatic analysis. Z.Y. designed experiments, supervised the project and wrote the paper.

Corresponding author

Correspondence to Zhen Yan.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Additional information

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

Integrated supplementary information

Supplementary Figure 1 Romidepsin treatment leads to the sustained increase of social interaction time and social preference in young Shank3-deficient mice.

(a) Box plots showing the time spent investigating either the social (Soc) or nonsocial (NS) stimulus during sociability testing in young (5-6 weeks old) male Shank3+/ΔC mice with a brief treatment of romidepsin (RMD, 0.25 mg/kg, i.p., 3x, n=10) or saline (n=10) prior to and at different days after the injection. F3,36(treatment)=137.4, P<0.0001; +++ P<0.001, (Soc vs. NS), ** P<0.01, *** P<0.001 (saline vs. romidepsin), two-way rmANOVA. (b) Representative heat maps illustrating the time spent in different locations of the 3 chambers from the social preference tests in Shank3+/ΔC mice treated with RMD or saline. Locations of Soc and NS stimuli are labeled with the circles. (c) Bar graphs (mean ± SEM) and scatter plots showing the preference index of the sociability testing in adult (10-11 weeks old) Shank3+/ΔC (n=13) or WT (n=10) mice before and after romidepsin (0.25 mg/kg, i.p., 3x) treatment. F3,45=10.5, P<0.0001; * P<0.05,*** P<0.001, one-way ANOVA. (d) Bar graphs (mean ± SEM) and scatter plots showing the social preference index in Shank3+/ΔC mice (n=9, ~10 weeks old) before and after the second-round romidepsin treatment. F2,24=2.6, ^ P=0.09, one-way ANOVA.

Supplementary Figure 2 Current antipsychotics fail to increase social interaction time, while the pan-HDAC inhibitor trichostatin A (TSA) transiently improves social behaviors in Shank3-deficient mice.

(ae) Box plots showing the time spent investigating either the social (Soc) or nonsocial (NS) stimulus during sociability testing in young (5-6 weeks old) male Shank3+/ΔC mice treated with fluoxetine (10 mg/kg, i.p., 14x, a, n=9), clozapine (5 mg/kg, i.p., 3x, b, n=11), valproic acid (VPA, 100 mg/kg, 3x, c, n=11), aripiprazole (1 mg/kg, 3x, d, n=9) or risperidone (0.1 mg/kg, 3x, e, n=10). In (c), F2,40(treatment)=0.71, P=0.5; +++ P<0.001 (Soc vs. NS), two-way rmANOVA. (f, g) Plots showing the preference index (f) and the time spent investigating either the Soc or NS stimulus (g) during sociability testing in Shank3+/ΔC mice before and after TSA treatment (0.5 mg/kg, i.p., 3x, n=8). In (f), F2,21=19.7, P<0.0001, one-way ANOVA. In (g), F2,28(treatment)=4.15, P=0.026; ++ P<0.01, +++ P<0.001 (Soc vs. NS), ### P<0.001 (pre- vs. post-injection), two-way rmANOVA. Inset (d,e,g): Representative heat maps illustrating the time spent in different locations of the 3 chambers (blue: 0 sec; red: ~20 sec) from the social preference tests of drug-treated Shank3+/ΔC mice.

Supplementary Figure 3 Histone acetylation sites are identified on Grin2a and Grin2b promoters.

(a) PCR images showing the ChIP (AcetyH3-occupied DNA), input (total DNA) and no-template control (NTC) signals with 3 primers (P1, P2, P3) designed against the promoter regions of Grin2a and Grin2b. The expected sizes of PCR products are labeled on the gels. The final primers used in the quantitative ChIP experiments are labeled by red circles. Top: diagram showing the primer locations on the 5′ upstream sequence of Grin2a and Grin2b. TSS, transcriptional start site. (b) The amplification curve and melt curve for AcetyH3-ChIP, input, and IgG control samples.

Supplementary Figure 4 Romidepsin treatment restores NMDAR synaptic function and global histone acetylation at 16–18 d, but not 30–32 d, postinjection.

(a,b) Input-output curves of NMDAR-EPSC in PFC pyramidal neurons from Shank3+/ΔC mice (Het, male) receiving treatment of romidepsin (RMD, i.p., 0.25 mg/kg, 3x) or saline. Recordings were performed at 16-18 days (a) or 30-32 days (b) post-injection. n=12 cells/3 mice each group. In (a), F1,22(treatment)=13.56, P=0.0013; * P<0.05, *** P<0.001, two-way rmANOVA. Data are mean ± SEM. Inset: representative NMDAR-EPSC traces. (c,d) Immunoblots and quantification analysis of the level of acetylated H3 and total H3 in the nuclear fraction of cortical slices from WT or Shank3+/ΔC mice (male) treated with saline or romidepsin at 16-18 days or 30-32 days post-injection. n=6 each group. In (d), F2,15=12.55, P=0.0006 (16-18 days); F2,15=27.72, P<0.0001 (30-32 days); ** P<0.01, *** P<0.001, ns, not significant, one-way ANOVA.

Supplementary Figure 5 Histone acetylation sites are identified on Arhgef7 and Limk1 promoters.

(a,b) PCR images showing the ChIP (AcetyH3-occupied DNA), input (total DNA) and no-template control (NTC) signals with 3 primers (P1, P2, P3) designed against the promoter regions of Arhgef7 and Limk1. The expected sizes of PCR products are labeled on the gels. Top Inset: diagram showing the primer locations on the 5′ upstream sequence of Arhgef7 and Limk1. TSS, transcriptional start site. Right Inset: The amplification curve and melt curve for AcetyH3-ChIP, input, and NTC samples. The final primers used in the quantitative ChIP experiments are labeled by red circles.

Supplementary Figure 6 Romidepsin treatment restores actin filaments in PFC of Shank3-deficient mice.

(a) High magnification confocal images (40x) of F-actin staining with phalloidin (co-stained with PSD-95 and DAPI) in PFC slices of WT vs. Shank3+/ΔC mice (5-6 weeks old, male) with i.p. injections of saline or romidepsin (0.25 mg/kg, 3x). (b) Quantification of PSD-95 levels (integrated densities) in PFC slices of different animal groups. n=27 images/3 mice each group.

Supplementary Figure 7 Romidepsin treatment has no effect on the expression of genes encoding NMDAR subunits or actin regulators in other brain regions and peripheral organs of Shank3-deficient mice.

(a,b) Quantitative real-time RT-PCR data on the mRNA level of Grin1, Grin2a, Grin2b, Arhgef7 and Limk1 in striatum, VTA, kidney and heart from WT or Shank3+/ΔC mice (5-6 weeks old, male) injected with saline or romidepsin (0.25 mg/kg, i.p., 3x). n=6 each group.

Supplementary Figure 8

Full blots for Figs. 1, 3 and 4.

Supplementary Figure 9

Full blots for Figs. 5 and 6 and Supplementary Figure 4.

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Qin, L., Ma, K., Wang, ZJ. et al. Social deficits in Shank3-deficient mouse models of autism are rescued by histone deacetylase (HDAC) inhibition. Nat Neurosci 21, 564–575 (2018). https://doi.org/10.1038/s41593-018-0110-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-018-0110-8

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