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:

Histone deacetylase 3 associates with MeCP2 to regulate FOXO and social behavior

Abstract

Mutations in MECP2 cause the neurodevelopmental disorder Rett syndrome (RTT). The RTT missense MECP2R306C mutation prevents MeCP2 from interacting with the NCoR/histone deacetylase 3 (HDAC3) complex; however, the neuronal function of HDAC3 is incompletely understood. We found that neuronal deletion of Hdac3 in mice elicited abnormal locomotor coordination, sociability and cognition. Transcriptional and chromatin profiling revealed that HDAC3 positively regulated a subset of genes and was recruited to active gene promoters via MeCP2. HDAC3-associated promoters were enriched for the FOXO transcription factors, and FOXO acetylation was elevated in Hdac3 knockout (KO) and Mecp2 KO neurons. Human RTT-patient-derived MECP2R306C neural progenitor cells had deficits in HDAC3 and FOXO recruitment and gene expression. Gene editing of MECP2R306C cells to generate isogenic controls rescued HDAC3-FOXO-mediated impairments in gene expression. Our data suggest that HDAC3 interaction with MeCP2 positively regulates a subset of neuronal genes through FOXO deacetylation, and disruption of HDAC3 contributes to cognitive and social impairment.

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: Hdac3 cKO mice exhibit social and cognitive impairments similar to models of RTT.
Figure 2: HDAC3 regulates the transcription of a subset of genes in common with MeCP2.
Figure 3: MeCP2 modulates HDAC3 localization to the promoters of transcribed genes.
Figure 4: HDAC3 regulates FOXO3 deacetylation and recruitment to gene promoters in neurons.
Figure 5: FOXO3 and HDAC3 recruitment to gene regulatory regions is abrogated in NPCs harboring a patient-derived MECP2 mutation.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

References

  1. Pohodich, A.E. & Zoghbi, H.Y. Rett syndrome: disruption of epigenetic control of postnatal neurological functions. Hum. Mol. Genet. 24, R10–R16 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Jones, P.L. et al. Methylated DNA and MeCP2 recruit histone deacetylase to repress transcription. Nat. Genet. 19, 187–191 (1998).

    CAS  PubMed  Google Scholar 

  3. Nan, X. et al. Transcriptional repression by the methyl-CpG-binding protein MeCP2 involves a histone deacetylase complex. Nature 393, 386–389 (1998).

    CAS  PubMed  Google Scholar 

  4. Mellén, M., Ayata, P., Dewell, S., Kriaucionis, S. & Heintz, N. MeCP2 binds to 5hmC enriched within active genes and accessible chromatin in the nervous system. Cell 151, 1417–1430 (2012).

    PubMed  PubMed Central  Google Scholar 

  5. Chahrour, M. et al. MeCP2, a key contributor to neurological disease, activates and represses transcription. Science 320, 1224–1229 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Ben-Shachar, S., Chahrour, M., Thaller, C., Shaw, C.A. & Zoghbi, H.Y. Mouse models of MeCP2 disorders share gene expression changes in the cerebellum and hypothalamus. Hum. Mol. Genet. 18, 2431–2442 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Li, Y. et al. Global transcriptional and translational repression in human-embryonic-stem-cell-derived Rett syndrome neurons. Cell Stem Cell 13, 446–458 (2013).

    PubMed  PubMed Central  Google Scholar 

  8. Yasui, D.H. et al. Integrated epigenomic analyses of neuronal MeCP2 reveal a role for long-range interaction with active genes. Proc. Natl. Acad. Sci. USA 104, 19416–19421 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Wang, Z. et al. Genome-wide mapping of HATs and HDACs reveals distinct functions in active and inactive genes. Cell 138, 1019–1031 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Cheng, J. et al. A role for H3K4 monomethylation in gene repression and partitioning of chromatin readers. Mol. Cell 53, 979–992 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Lopez-Atalaya, J.P., Ito, S., Valor, L.M., Benito, E. & Barco, A. Genomic targets, and histone acetylation and gene expression profiling of neural HDAC inhibition. Nucleic Acids Res. 41, 8072–8084 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Zupkovitz, G. et al. Negative and positive regulation of gene expression by mouse histone deacetylase 1. Mol. Cell. Biol. 26, 7913–7928 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Lyst, M.J. et al. Rett syndrome mutations abolish the interaction of MeCP2 with the NCoR/SMRT co-repressor. Nat. Neurosci. 16, 898–902 (2013).

    CAS  PubMed  Google Scholar 

  14. Ebert, D.H. et al. Activity-dependent phosphorylation of MeCP2 threonine 308 regulates interaction with NCoR. Nature 499, 341–345 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Norwood, J., Franklin, J.M., Sharma, D. & D'Mello, S.R. Histone deacetylase 3 is necessary for proper brain development. J. Biol. Chem. 289, 34569–34582 (2014).

    PubMed  PubMed Central  Google Scholar 

  16. Montgomery, R.L. et al. Maintenance of cardiac energy metabolism by histone deacetylase 3 in mice. J. Clin. Invest. 118, 3588–3597 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Zeng, H. et al. Forebrain-specific calcineurin knockout selectively impairs bidirectional synaptic plasticity and working/episodic-like memory. Cell 107, 617–629 (2001).

    CAS  PubMed  Google Scholar 

  18. Gemelli, T. et al. Postnatal loss of methyl-CpG binding protein 2 in the forebrain is sufficient to mediate behavioral aspects of Rett syndrome in mice. Biol. Psychiatry 59, 468–476 (2006).

    CAS  PubMed  Google Scholar 

  19. Su, S.-H., Kao, F.-C., Huang, Y.-B. & Liao, W. MeCP2 in the rostral striatum maintains local dopamine content critical for psychomotor control. J. Neurosci. 35, 6209–6220 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Guy, J., Hendrich, B., Holmes, M., Martin, J.E. & Bird, A. A mouse Mecp2-null mutation causes neurological symptoms that mimic Rett syndrome. Nat. Genet. 27, 322–326 (2001).

    CAS  PubMed  Google Scholar 

  21. Moretti, P., Bouwknecht, J.A., Teague, R., Paylor, R. & Zoghbi, H.Y. Abnormalities of social interactions and home-cage behavior in a mouse model of Rett syndrome. Hum. Mol. Genet. 14, 205–220 (2005).

    CAS  PubMed  Google Scholar 

  22. Yasui, D.H. et al. Mice with an isoform-ablating Mecp2 exon 1 mutation recapitulate the neurologic deficits of Rett syndrome. Hum. Mol. Genet. 23, 2447–2458 (2014).

    CAS  PubMed  Google Scholar 

  23. Pelka, G.J. et al. Mecp2 deficiency is associated with learning and cognitive deficits and altered gene activity in the hippocampal region of mice. Brain 129, 887–898 (2006).

    PubMed  Google Scholar 

  24. Moretti, P. et al. Learning and memory and synaptic plasticity are impaired in a mouse model of Rett syndrome. J. Neurosci. 26, 319–327 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Miyakawa, T., Yamada, M., Duttaroy, A. & Wess, J. Hyperactivity and intact hippocampus-dependent learning in mice lacking the M1 muscarinic acetylcholine receptor. J. Neurosci. 21, 5239–5250 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Frankland, P.W., O'Brien, C., Ohno, M., Kirkwood, A. & Silva, A.J. Alpha-CaMKII-dependent plasticity in the cortex is required for permanent memory. Nature 411, 309–313 (2001).

    CAS  PubMed  Google Scholar 

  27. Baker, S.A. et al. An AT-hook domain in MeCP2 determines the clinical course of Rett syndrome and related disorders. Cell 152, 984–996 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Kumar, V. et al. Uniform, optimal signal processing of mapped deep-sequencing data. Nat. Biotechnol. 31, 615–622 (2013).

    CAS  PubMed  Google Scholar 

  29. Ernst, J. & Kellis, M. Discovery and characterization of chromatin states for systematic annotation of the human genome. Nat. Biotechnol. 28, 817–825 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Gjoneska, E. et al. Conserved epigenomic signals in mice and humans reveal immune basis of Alzheimer's disease. Nature 518, 365–369 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Peleg, S. et al. Altered histone acetylation is associated with age-dependent memory impairment in mice. Science 328, 753–756 (2010).

    CAS  PubMed  Google Scholar 

  32. Bhaskara, S. et al. Deletion of histone deacetylase 3 reveals critical roles in S phase progression and DNA damage control. Mol. Cell 30, 61–72 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Bhaskara, S. et al. Hdac3 is essential for the maintenance of chromatin structure and genome stability. Cancer Cell 18, 436–447 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Mihaylova, M.M. et al. Class IIa histone deacetylases are hormone-activated regulators of FOXO and mammalian glucose homeostasis. Cell 145, 607–621 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Daitoku, H. et al. Silent information regulator 2 potentiates Foxo1-mediated transcription through its deacetylase activity. Proc. Natl. Acad. Sci. USA 101, 10042–10047 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Hatta, M., Liu, F. & Cirillo, L.A. Acetylation curtails nucleosome binding, not stable nucleosome remodeling, by FoxO1. Biochem. Biophys. Res. Commun. 379, 1005–1008 (2009).

    CAS  PubMed  Google Scholar 

  37. Matsuzaki, H. et al. Acetylation of Foxo1 alters its DNA-binding ability and sensitivity to phosphorylation. Proc. Natl. Acad. Sci. USA 102, 11278–11283 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Cheung, A.Y.L. et al. Isolation of MECP2-null Rett Syndrome patient hiPS cells and isogenic controls through X-chromosome inactivation. Hum. Mol. Genet. 20, 2103–2115 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Wen, Z. et al. Synaptic dysregulation in a human iPS cell model of mental disorders. Nature 515, 414–418 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. McQuown, S.C. et al. HDAC3 is a critical negative regulator of long-term memory formation. J. Neurosci. 31, 764–774 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Gabel, H.W. et al. Disruption of DNA-methylation-dependent long gene repression in Rett syndrome. Nature 522, 89–93 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Kinde, B., Gabel, H.W., Gilbert, C.S., Griffith, E.C. & Greenberg, M.E. Reading the unique DNA methylation landscape of the brain: non-CpG methylation, hydroxymethylation, and MeCP2. Proc. Natl. Acad. Sci. USA 112, 6800–6806 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Choudhary, C. et al. Lysine acetylation targets protein complexes and co-regulates major cellular functions. Science 325, 834–840 (2009).

    CAS  PubMed  Google Scholar 

  44. Krumm, N., O'Roak, B.J., Shendure, J. & Eichler, E.E. A de novo convergence of autism genetics and molecular neuroscience. Trends Neurosci. 37, 95–105 (2014).

    CAS  PubMed  Google Scholar 

  45. Hormozdiari, F., Penn, O., Borenstein, E. & Eichler, E. The discovery of integrated gene networks for autism and related disorders. Genome Res. 25, 142–154 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. O'Roak, B.J. et al. Multiplex targeted sequencing identifies recurrently mutated genes in autism spectrum disorders. Science 338, 1619–1622 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Chung, R.-H. et al. An X chromosome-wide association study in autism families identifies TBL1X as a novel autism spectrum disorder candidate gene in males. Mol. Autism 2, 18 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Pons, L. et al. A new syndrome of intellectual disability with dysmorphism due to TBL1XR1 deletion. Am. J. Med. Genet. A. 167A, 164–168 (2015).

    PubMed  Google Scholar 

  49. Saitsu, H. et al. A girl with West syndrome and autistic features harboring a de novo TBL1XR1 mutation. J. Hum. Genet. 59, 581–583 (2014).

    CAS  PubMed  Google Scholar 

  50. Shi, Y., Kirwan, P. & Livesey, F.J. Directed differentiation of human pluripotent stem cells to cerebral cortex neurons and neural networks. Nat. Protoc. 7, 1836–1846 (2012).

    CAS  PubMed  Google Scholar 

  51. Anagnostaras, S.G., Josselyn, S.A., Frankland, P.W. & Silva, A.J. Computer-assisted behavioral assessment of Pavlovian fear conditioning in mice. Learn. Mem. 7, 58–72 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Livak, K.J. & Schmittgen, T.D. Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods 25, 402–408 (2001).

    CAS  PubMed  Google Scholar 

  53. Feng, D. et al. A circadian rhythm orchestrated by histone deacetylase 3 controls hepatic lipid metabolism. Science 331, 1315–1319 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Eijkelenboom, A. et al. Genome-wide analysis of FOXO3 mediated transcription regulation through RNA polymerase II profiling. Mol. Syst. Biol. 9, 638 (2013).

    PubMed  PubMed Central  Google Scholar 

  55. Karmodiya, K., Krebs, A.R., Oulad-Abdelghani, M., Kimura, H. & Tora, L. H3K9 and H3K14 acetylation co-occur at many gene regulatory elements, while H3K14ac marks a subset of inactive inducible promoters in mouse embryonic stem cells. BMC Genomics 13, 424 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Ernst, J. & Kellis, M. ChromHMM: automating chromatin-state discovery and characterization. Nat. Methods 9, 215–216 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank E.N. Olson (UT Southwestern Medical Center) for kindly providing the HDAC3f/f mice. We thank M. Sur and A. Banerjee (both at Massachusetts Institute of Technology for providing Mecp2 KO mice. We thank G.-L. Ming (Johns Hopkins University) for kindly providing the C1 iPSC cell line. We thank R. Madabhushi, A. Watson and J. Penney for comments on the manuscript. We thank E. Demmons for help with mouse colony maintenance. This work was supported by US National Institutes of Health grants (MH102690 and NS079625) and Rettsyndrome.org to P.J., and NIH grant NS78839 and the JPB Foundation to L.-H.T.

Author information

Authors and Affiliations

Authors

Contributions

A.N. and L.-H.T. designed the study, and L.-H.T. directed and coordinated the study. A.N. initiated, planned and performed experimental work. J.C. helped with ChIP experiments. F.G. and J.M. contributed to bioinformatic analysis. P.M. helped with some behavioral experiments. A.V.Z. helped with validation of the RNA-seq. E.G. prepared sequencing libraries for the ChIP-seq experiment. T.K. derived the NPC lines. T.K. and Y.-T.L. helped to maintain the iPSCs. F.Z. and P.J. provided Mecp2 KO tissue. A.N. and L.-H.T. wrote the manuscript with critical input from all of the authors.

Corresponding author

Correspondence to Alexi Nott.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Hdac3 forebrain-specific conditional knockout from excitatory neurons.

(a) Coronal brain slices of 3-month control and Hdac3 cKO male mice were stained using Hoechst (blue) and anti-HDAC3 (green), and images captured of the hippocampus using tile scan of a confocal microscope. Scale bar, 200 μm. (b) Mean gray value quantification of HDAC3 immunofluorescence was compared between control and Hdac3 cKO mice in the CA1 and dentate gyrus (DG) of the hippocampus; two-tailed t test; CA1, t(8) = 12.88, **** P < 0.0001; DG, t(8) = 5.958, *** P = 0.0003; n = 4, 6 mice. (c) Coronal brain slices of 3-month control and Hdac3 cKO male mice were stained using Hoechst (blue), anti-HDAC3 (red), anti-GFAP (green), and anti-NeuN (white), and images captured of hippocampal CA1 using a confocal microscope. Scale bar, 20 μm; n = 3, 3 mice. (d) Coronal brain slices of 3-month control and Hdac3 cKO male mice were stained using Hoechst (blue), anti-HDAC3 (red), anti-Parvalbumin (green), and anti-NeuN (white), and images captured of hippocampal CA1 using a confocal microscope. Scale bar, 20 μm; n = 3, 3 mice. (e) Western blot analysis of HDAC3, HDAC2 and actin protein levels using hippocampal whole cell lysates from control and Hdac3 cKO mice. Full-length WB’s are presented in Supplementary Fig. 10b. (f) Mean gray value quantification of HDAC3 protein levels normalized to actin of western blot in c. Two-tailed t test; t(6) = 9.663, **** P < 0.0001; n = 4, 4 mice. (g) Western blot analysis of HDAC3 and actin protein levels using cortex whole cell lysates from control and Hdac3 cKO mice. Full-length WB’s are presented in Supplementary Fig. 10c. (h) Mean gray value quantification of HDAC3 protein levels normalized to actin of western blot in e. Two-tailed t test; t(6) = 3.460, * P = 0.0135; n = 4, 4 mice. (i), Western blot analysis of HDAC3 and actin protein levels using striatum whole cell lysates from control and Hdac3 cKO mice. Full-length WB’s are presented in Supplementary Fig. 10d. (j) Mean gray value quantification of HDAC3 protein levels normalized to actin of western blot in g. Two-tailed t test; t(6) = 0.9516, P = 0.3780; n = 4, 4 mice. All values are mean ± s.e.m.

Supplementary Figure 2 Hdac3 cKO mice exhibit locomotor abnormalities.

(a) (f), Control and Hdac3 cKO mice were subjected to open-field test for 60 minutes and the data was analyzed in 5 minute bins using two-way ANOVA with Bonferroni post hoc. Two-way ANOVA for, (a) vertical activity; Genotype: F(1,44) = 150.1, **** P < 0.0001, Time: F(11, 484) = 18.45, **** P < 0.0001 (Bonferroni post hoc **** P < 0.0001), (b) total distance traveled; Genotype: F(1,44) = 57.18, **** P < 0.0001, Time: F(11, 484) = 47.40, **** P < 0.0001 (Bonferroni post hoc **** P < 0.0001), (c) horizontal activity; Genotype: F(1,44) = 136.26, **** P < 0.0001, Time: F(11, 484) = 38.26, **** P < 0.0001 (Bonferroni post hoc **** P < 0.0001), (d) margin distance traveled Genotype: F(1,44) = 95.66, **** P < 0.0001, Time: F(11, 484) = 63.44, **** P < 0.0001 (Bonferroni post hoc **** P < 0.0001), (e) center distance traveled; Genotype: F(1,44) = 72.76, **** P < 0.0001, Time: F(11, 484) = 5.077, **** P < 0.0001 (Bonferroni post hoc **** P < 0.0001, * P < 0.05), and (f) time spent in the margin; Genotype: F(1,44) = 24.82, **** P < 0.0001, Time: F(11, 484) = 2.131, * P = 0.0171 (Bonferroni post hoc **** P < 0.0001, * P < 0.05); for a-f, n = 22, 24 mice. All values are mean ± s.e.m.

Supplementary Figure 3 Hdac3 cKO mice exhibit deficits in sociability and hippocampal-dependent learning.

(a) Time spent in chambers during 10 minute habituation phase of the sociability task (Fig. 1d) for control and Hdac3 cKO mice. Control; one-way ANOVA, F(2, 24) = 0.1687, P = 0.8457; Bonferroni post hoc, Left versus Center, P > 0.99; Left versus Right, P > 0.99; Center versus Right, P > 0.99; n = 9 mice. Hdac3 cKO; one-way ANOVA, F(2, 24) = 2.793, P = 0.0812; Bonferroni post hoc, Left versus Center, P = 0.4200; Left versus Right, P > 0.99; Center versus Right, P = 0.0864; n = 9, 9 mice. (b) Total time spent with both objects (s) during the object location memory task (Fig. 1e); two-tailed t test; t(26) = 3.883; *** P = 0.0006; n = 16, 12 mice. (c) The average swim speed (m s−1) of the mice was recorded during the training days (day 1 – 7) and on the probe trial day (day 8) of the MWM task. Two-way ANOVA; Genotype, F(1,133) = 9.530, ** P =0.0025; Day, F(6,133) = 3.050, ** P = 0.0079; Bonferroni post hoc; day 1, P > 0.99; day 2, P > 0.99; day 3, P > 0.99; day 4, P > 0.99; day 5, P = 0.0677; day 6, P > 0.99; day 7, P = 0.2703; n = 10, 11 mice. (d) Mean velocity (cm s−1) measured during the habituation and the e-stimulation phases on the training day of fear-conditioned learning; Control: two-tailed t test; t(20) = 15.13; **** P < 0.0001; Hdac3 cKO: t(20) = 17.92; **** P < 0.0001; n = 11, 11 mice. (e) Activity suppression of control and Hdac3 cKO mice during testing of contextual memory; two-tailed t test; t(20) = 3.973; *** P = 0.0007; n = 11, 11 mice. (f) Activity suppression of control and Hdac3 cKO mice during testing of cued (tone) memory; two-tailed t test; t(20) = 5.069; **** P < 0.0001; n = 11, 11 mice. c, report mean value ± s.e.m. a, b, d - f, report median, 25th and 75th percentile, min and max value.

Supplementary Figure 4 Transcriptional regulation in the Hdac3 cKO.

(a) qRT-PCR analysis of RNA extracted from the CA1 region of control and Hdac3 cKO mice. Primers designed within exon 10 (non-floxed) and exon 11 (floxed). Two-tailed t test; t(6) = 7.390, *** P = 0.0003; n = 4, 4 mice. Values are mean ± s.e.m. (b) RT-PCR analysis of RNA extracted from the CA1 region of control and Hdac3 cKO mice. Primers designed within exons 10 and 15, flanking floxed exons 11 – 14, n = 2, 2 mice. (c) log2fold change analysis of Hdac3 exon read counts between control and Hdac3 cKO mice following RNA-sequencing of the CA1 region of the hippocampus; n = 2, 2 mice. (d) UCSC genome browser visualization of RNA-seq coverage at the Hdac3 locus for control (blue) and Hdac3 cKO (red). The y-axis is number of reads mapped.

Supplementary Figure 5 HDAC3 is recruited to promoters of transcribed genes.

(a) Enrichment of HDAC3 binding peaks at annotated regions of the genome analyzed using Homer (enrichment over genome). TTS, transcriptional termination site; UTR, untranslated region. (b) Enrichment of HDAC3 binding peaks at genomic regions analyzed using chromatin state profiling of the hippocampus. TSS, transcriptional start site. (c) Aggregate plots of average HDAC3 ChIP-seq intensity signal in 3-month wild-type hippocampus. Aggregate plots were generated at the TSS of genes upregulated (green) and downregulated (blue) in the CA1 of Hdac3 cKO mice. HDAC3 ChIP-seq intensity signal at the TSS of genes with FPKM expression values below 0.1 (black; negative control). TSS, transcriptional start site. (d) UCSC genome browser view of Tle1 and Klf10 loci showing RNA-seq coverage from control and Hdac3 cKO hippocampus, and HDAC3 ChIP-seq coverage and peak calling from wild-type hippocampus. (e) ChIP of HDAC3 and IgG followed by qPCR analysis at the promoters of Arrdc2, Dusp4, Klf10, Tle1, Bdnf, and Nr4a1 in hippocampus of wild-type mice (% input); two-tailed t test; Arrdc2, t(4) = 6.944, ** P = 0.0023; Dusp4, t(4) = 5.074, ** P = 0.0071; Klf10, t(4) = 7.636, ** P = 0.0016; Tle1, t(4) = 2.917, * P = 0.0434; Bdnf, t(4) = 3.350, * P = 0.0286; Nr4a1, t(3) = 6.237, ** P = 0.0083; n = 3, 3 mice. (f) ChIP of HDAC3 and IgG followed by qPCR analysis at the promoters of Arrdc2, Dusp4, Klf10, Tle1, Bdnf, and Nr4a1 in hippocampal CA1 region of control and Hdac3 cKO mice (% input); One-way ANOVA; Arrdc2, F(2,6) = 9.653, * P = 0.0133; Dusp4, F(2,6) = 20.38, ** P = 0.0021; Klf10, F(2,6) = 77.61, **** P < 0.0001; Tle1, F(2,6) = 19.07, ** P = 0.0025; Bdnf, F(2,6) = 15.94, ** P = 0.0040; Nr4a1, F(2,6) = 54.27, *** P = 0.0001. Bonferroni post hoc; control versus Hdac3 KO: Arrdc2, * P = 0.0472; Dusp4, * P = 0.0246; Klf10, *** P = 0.0006; Tle1, * P = 0.0368; Bdnf, * P = 0.0201; Nr4a1, *** P = 0.0006; control versus IgG: Arrdc2, * P = 0.0105; Dusp4, ** P = 0.0014; Klf10, **** P < 0.0001; Tle1, ** P = 0.0017; Bdnf, ** P = 0.0029; Nr4a1, *** P = 0.0001; n = 3, 3, 3 mice. e, f, report mean value ± s.e.m.

Supplementary Figure 6 HDAC3 and MeCP2 regulate the recruitment of FOXO transcription factors to gene promoters.

(a) UCSC genome browser view at the Arrdc2 locus of HDAC3 ChIP-seq coverage in the hippocampus of P45 wild-type and Mecp2 KO mice. (b) Aggregate plots of average HDAC3 ChIP-seq signal in P45 wild-type hippocampus (blue) and Mecp2 KO hippocampus (purple). Aggregate plots were generated at the TSS of genes upregulated in the hippocampus of Mecp2 KO mice27. HDAC3 ChIP-seq signal at the TSS of genes with the lowest expression in P45 hippocampus27 (black; negative control; 10% of total genes). TSS, transcriptional start site. (c) Aggregate plots of average HDAC3 ChIP-seq signal in P45 wild-type hippocampus (blue) and Mecp2 KO hippocampus (purple). Aggregate plots were generated at the TSS of genes downregulated in the hippocampus of Mecp2 KO mice. HDAC3 ChIP-seq signal at the TSS of genes with the lowest expression in P45 hippocampus27 (black; negative control; 10% of total genes). TSS, transcriptional start site. (d) ChIP of MeCP2 and IgG followed by qPCR at the promoters of Arrdc2, Dusp4, Klf10, Tle1, Bdnf and Nr4a1 in hippocampus of wildt-ype and Mecp2 KO mice (% input). One-way ANOVA; Arrdc2, F(2,9) = 12.62, ** P = 0.0024; Dusp4, F(2,9) = 28.96, **** P < 0.0001; Klf10, F(2,9) = 14.84, ** P = 0.0014; Tle1, F(2,9) = 14.39, ** P = 0.0016; Bdnf, F(2,9) = 8.119, ** P = 0.0097; Nr4a1, F(2,9) = 5.996, * P = 0.0221. Bonferroni post hoc; wild-type versus Mecp2 KO: Arrdc2, ** P = 0.0028; Dusp4, *** P = 0.0003; Klf10, ** P = 0.0053; Tle1, ** P = 0.0037; Bdnf, * P = 0.0307; Nr4a1, * P = 0.0231; wild-type versus IgG: Arrdc2, ** P = 0.0051; Dusp4, *** P = 0.0001; Klf10, ** P = 0.0012; Tle1, ** P = 0.0017; Bdnf, ** P = 0.0080; Nr4a1, * P = 0.0410; n = 4, 4, 4 mice. (e) ChIP of H4K12ac and IgG followed by qPCR analysis at the promoters of Arrdc2, Dusp4, Klf10, Tle1, Bdnf, and Nr4a1 in hippocampal CA1 region of control and Hdac3 cKO mice (% input); one-way ANOVA; Arrdc2, F(2,6) = 18.79, ** P = 0.0026; Dusp4, F(2,6) = 7.802, * P = 0.0214; Klf10, F(2,6) = 10.09, ** P = 0.0120; Tle1, F(2,6) = 6.464, * P = 0.0318; Bdnf, F(2,6) = 8.743, * P = 0.0167; Nr4a1, F(2,6) = 5.209, * P = 0.0488. Bonferroni post hoc; Control versus Hdac3 cKO: Arrdc2, P = 0.5511; Dusp4, P > 0.99; Klf10, P = 0.4278; Tle1, P > 0.99; Bdnf, P > 0.99; Nr4a1, P > 0.99; Control versus IgG: Arrdc2, ** P = 0.0023; Dusp4, * P = 0.0168; Klf10, * P = 0.0478; Tle1, * P = 0.0310; Bdnf, * P = 0.0245; Nr4a1, * P = 0.0448; n = 3, 3, 3 mice. (f) ChIP of H3K9ac and IgG followed by qPCR analysis at the promoters of Arrdc2, Dusp4, Klf10, Tle1, Bdnf, and Nr4a1 in hippocampal CA1 region of control and Hdac3 cKO mice (% input); one-way ANOVA; Arrdc2, F(2,5) = 7.810, * P = 0.0290; Dusp4, F(2,5) = 18.34, ** P = 0.0050; Klf10, F(2,5) = 8.792, * P = 0.0231; Tle1, F(2,5) = 9.315, * P = 0.0206; Bdnf, F(2,5) = 43.53, * P = 0.0007; Nr4a1, F(2,5) = 30.06, ** P = 0.0016. Bonferroni post hoc; Control versus Hdac3 cKO: Arrdc2, P > 0.99; Dusp4, P > 0.99; Klf10, P > 0.99; Tle1, P > 0.99; Bdnf, P = 0.8642; Nr4a1, P = 0.9790; Control versus IgG: Arrdc2, * P = 0.0317; Dusp4, ** P = 0.0058; Klf10, * P = 0.0248; Tle1, * P = 0.0318; Bdnf, *** P = 0.0007; Nr4a1, ** P = 0.0016; n = 3, 3, 2 mice. All values are mean ± s.e.m.

Supplementary Figure 7 HDAC3 and MeCP2 regulate the recruitment of FOXO transcription factors to gene promoters.

(a) Upper table: transcription factor-binding motifs enriched at the 194 genes downregulated in Hdac3 cKO mice. Lower table: the FOXO motif (TTGTTT) was enriched within the 1274 genes downregulated in the hippocampus of Mecp2 KO mice27, the 2167 genes downregulated in the hypothalamus of Mecp2 KO mice5, and the 410 genes downregulated in the cerebellum of Mecp2 KO mice6. (b) Frequency probability matrix for de novo motif identified by MEME-ChIP analysis of HDAC3 binding regions; input of 6149 HDAC3 binding sequences of which 1038 sites identified with TRTTTW motif. (c) Foxo1, Foxo3, Foxo4 and Foxo6 gene expression in the CA1 hippocampus of control and Hdac3 cKO mice; n = 2, 2 mice. FPKM, fragments per kilobase of exon per million fragments mapped. (d) Coronal brain slices of control and Hdac3 cKO mice were stained using Hoechst (blue), anti-acetyl-FOXO (green), and anti-NeuN (white). Images of the striatum using a confocal microscope, scale bar, 20 μm. Mean gray value of acetyl-FOXO immunofluorescence quantitated using only NeuN-positive nuclei in 3-month control and Hdac3 cKO mice. Two-tailed t test; t(10) = 0.9065, P = 0.3860; n = 5, 7 mice. (e) Immunoprecipitation of recombinant human FOXO3 with human MeCP2 or human HDAC3 followed by western blot analysis of FOXO3, HDAC3 and MeCP2; n = 3 assays. Full-length blots are presented in Supplementary Fig. 10e. (f) Immunoprecipitation of recombinant human MeCP2 with human HDAC3 followed by western blot analysis of MeCP2 and HDAC3; n = 3 assays. Full-length blots are presented in Supplementary Fig. 10f. (g) Immunostaining of C1 (control), line 4 and line 14 (MECP2R306C), line 6 and line 36 (isogenic control) NPCs using Hoechst (blue), anti-MSI1 (green), and anti-Nestin (red). Imaged using confocal microscope; lines C1, 4 and 14, n = 6, 6, 6 coverslips, 2 experiments; lines 6 and 36, n = 3, 3 coverslips, 1 experiment; scale bar = 40 μm. All values are mean ± s.e.m.

Supplementary Figure 8 Genotyping of iPSCs and NPCs derived from a MECP2R306C RTT patient and corresponding isogenic control lines generated by CRISPR/Cas9 gene editing.

(a) - (d) Chromatogram view of MECP2R306C locus following sequencing of genomic DNA (gDNA) and complementary DNA (cDNA) from iPSC and NPC lines as indicated. Sequences were analyzed for MECP2R306C iPSCs (a, b), line 4 and 14 MECP2R306C NPCs (c), line 6 and 36 iPSCs, and line 6 and 36 NPCs (d). Purple inset indicates the nucleotide position of the R306C mutation. Green inset indicates the nucleotide position of the corrected R306C mutation following CRISPR/Cas9 gene editing. Red insets indicate intentional introduction of heterozygous silent mutations following CRISPR/Cas9 gene editing.

Supplementary Figure 9 Chromatogram view of the five genomic regions with the highest probability for off-targeted CRISPR/Cas9-mediated gene editing.

Chromatograms show sequencing of gDNA from isogenic control lines 6 and 36. Off-target CRISPR/Cas9 genomic editing was not detected in either line. Black inset indicates the nucleotide region predicted for potential off-target gene editing.

Supplementary Figure 10 Full-length Western Blots

(a) Full-length western blots of Snap25 and Actin for cropped images in Fig. 2e. (b) Full-length western blots of HDAC3, HDAC2 and Actin for cropped images in Supplementary Fig. 1e. (c) Full-length western blots of HDAC3 and Actin for cropped images in Supplementary Fig. 1g. (d) Full-length western blots of HDAC3 and Actin for cropped images in Supplementary Fig. 1i. (e) Full-length western blots of HDAC3, FOXO3 and MeCP2 for cropped images in Supplementary Fig. 7e. (f) Full-length western blots of HDAC3 and MeCP2 for cropped images in Supplementary Fig. 7f.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–10 and Supplementary Tables 1–4 (PDF 4143 kb)

Supplementary Methods Checklist (PDF 551 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Nott, A., Cheng, J., Gao, F. et al. Histone deacetylase 3 associates with MeCP2 to regulate FOXO and social behavior. Nat Neurosci 19, 1497–1505 (2016). https://doi.org/10.1038/nn.4347

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4347

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