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.

  • Review Article
  • Published:

From neural development to cognition: unexpected roles for chromatin

An Erratum to this article was published on 09 May 2013

This article has been updated

Key Points

  • ATP-dependent chromatin remodellers and chromatin modifiers can have important roles in specific biological processes, such as brain development and function. Disruption of chromatin regulation can lead to neurodevelopmental and psychiatric diseases.

  • Recent exome-sequencing studies of human mental disorders have identified causal mutations in chromatin regulators. Newly uncovered mutations will guide researchers to important mechanistic studies of human mental disorders in the hope of identifying targets for disease treatments and learning more about the development and function of the brain.

  • So far, mutations in nine genes encoding seven subunits of SWI/SNF-like BAF chromatin-remodelling complexes have been identified in various mental disorders, including: intellectual disability, Coffin–Siris syndrome (CSS), Nicolaides–Baraitser syndrome (NBS), Kleefstra's syndrome, non-familial autism spectrum disorder (ASD) and schizophrenia. The essential switch of BAF complex subunits during differentiation of neural progenitors to postmitotic neurons as well as the rate-limiting role of BAF complexes in induced neuron formation indicate that neuronal dendritic morphogenesis and targeting might be a common underlying disease mechanism.

  • Recent exome-sequencing studies identified de novo mutations in multiple members of CHD family chromatin remodellers, most notably CHD8, in patients with non-familial autism spectrum disorder (ASD). We speculate that the interplay between CHD8 and the WNT/β-catenin signalling pathway, a major regulator of neurodevelopment, contributes to ASD phenotypes.

  • The enhancer of zeste 2 (EZH2) enzyme, a subunit of the Polycomb repressive complex 2 (PRC2), is mutated in numerous cases of Weaver's syndrome. EZH2 has a role in various neural development processes, and its antagonistic role to proteins, such as BAF complexes, on key neural development genes (for example, β-catenin targets) may be important for understanding disease mechanisms.

  • Mouse genetic studies showed that histone deacetylase 4 (HDAC4) is involved in learning, memory and long-term synaptic plasticity. This role is potentially the cause of brachydactyly mental retardation syndrome in patients with HDAC4 mutations.

  • Interestingly, many of the chromatin regulators mutated in human neurodevelopmental disorders are also mutated in human cancer. This may suggest unanticipated mechanisms of chromatin regulation in human diseases.

Abstract

Recent genome-sequencing studies in human neurodevelopmental and psychiatric disorders have uncovered mutations in many chromatin regulators. These human genetic studies, along with studies in model organisms, are providing insight into chromatin regulatory mechanisms in neural development and how alterations to these mechanisms can cause cognitive deficits, such as intellectual disability. We discuss several implicated chromatin regulators, including BAF (also known as SWI/SNF) and CHD8 chromatin remodellers, HDAC4 and the Polycomb component EZH2. Interestingly, mutations in EZH2 and certain BAF complex components have roles in both neurodevelopmental disorders and cancer, and overlapping point mutations are suggesting functionally important residues and domains. We speculate on the contribution of these similar mutations to disparate disorders.

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: Chromatin regulators have essential roles throughout neural development.
Figure 2: BAF complex roles in neurodevelopment and disorders of brain function.
Figure 3: Repressive chromatin modifiers involved in disorders of brain function.

Similar content being viewed by others

Change history

  • 09 May 2013

    In this article, some of the references citations were incorrectly numbered in the 'Chromatin remodellers' section and in Table 1. This has now been corrected online. The editors apologize for this mistake.

References

  1. Meissner, A. et al. Genome-scale DNA methylation maps of pluripotent and differentiated cells. Nature 454, 766–770 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Bernstein, B. E. et al. An integrated encyclopedia of DNA elements in the human genome. Nature 489, 57–74 (2012).

    Article  CAS  Google Scholar 

  3. Mikkelsen, T. S. et al. Genome-wide maps of chromatin state in pluripotent and lineage-committed cells. Nature 448, 553–560 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Hargreaves, D. C. & Crabtree, G. R. ATP-dependent chromatin remodeling: genetics, genomics and mechanisms. Cell Res. 21, 396–420 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Ho, L. & Crabtree, G. R. Chromatin remodelling during development. Nature 463, 474–484 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Jenuwein, T. & Allis, C. D. Translating the histone code. Science 293, 1074–1080 (2001).

    Article  CAS  PubMed  Google Scholar 

  7. Allis, C. D., Jenuwein, T. & Reinberg, D. Epigenetics (Cold Spring Harbor Laboratory Press, 2007).

    Google Scholar 

  8. Jones, P. A. Functions of DNA methylation: islands, start sites, gene bodies and beyond. Nature Rev. Genet. 13, 484–492 (2012).

    Article  CAS  PubMed  Google Scholar 

  9. Day, J. J. & Sweatt, J. D. Epigenetic mechanisms in cognition. Neuron 70, 813–829 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. van Bokhoven, H. Genetic and epigenetic networks in intellectual disabilities. Annu. Rev. Genet. 45, 81–104 (2011).

    Article  CAS  PubMed  Google Scholar 

  11. Jakovcevski, M. & Akbarian, S. Epigenetic mechanisms in neurological disease. Nature Med. 18, 1194–1204 (2012).

    Article  CAS  PubMed  Google Scholar 

  12. Mehler, M. F. Epigenetics and the nervous system. Ann. Neurol. 64, 602–617 (2008).

    Article  CAS  PubMed  Google Scholar 

  13. Yoo, A. S. & Crabtree, G. R. ATP-dependent chromatin remodeling in neural development. Curr. Opin. Neurobiol. 19, 120–126 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Corona, D. F. et al. ISWI regulates higher-order chromatin structure and histone H1 assembly in vivo. PLoS Biol. 5, e232 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Ho, L. et al. An embryonic stem cell chromatin remodeling complex, esBAF, is essential for embryonic stem cell self-renewal and pluripotency. Proc. Natl Acad. Sci. USA 106, 5181–5186 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Lessard, J. et al. An essential switch in subunit composition of a chromatin remodeling complex during neural development. Neuron 55, 201–215 (2007). This paper describes a neuron-specific chromatin-remodelling complex, nBAF, which has three subunits expressed only in the nervous system. The frequency with which BAF complexes are implicated in neurodevelopmental and psychiatric disorders might relate to the specialized functions of the nBAF complex during neural development.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Ooi, L., Belyaev, N. D., Miyake, K., Wood, I. C. & Buckley, N. J. BRG1 chromatin remodeling activity is required for efficient chromatin binding by repressor element 1-silencing transcription factor (REST) and facilitates REST-mediated repression. J. Biol. Chem. 281, 38974–38980 (2006).

    Article  CAS  PubMed  Google Scholar 

  18. Bultman, S. et al. A Brg1 null mutation in the mouse reveals functional differences among mammalian SWI/SNF complexes. Mol. Cell 6, 1287–1295 (2000).

    Article  CAS  PubMed  Google Scholar 

  19. Klochendler-Yeivin, A. et al. The murine SNF5/INI1 chromatin remodeling factor is essential for embryonic development and tumor suppression. EMBO Rep. 1, 500–506 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Kim, J. K. et al. Srg3, a mouse homolog of yeast SWI3, is essential for early embryogenesis and involved in brain development. Mol. Cell. Biol. 21, 7787–7795 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Wu, J. I. et al. Regulation of dendritic development by neuron-specific chromatin remodeling complexes. Neuron 56, 94–108 (2007).

    Article  CAS  PubMed  Google Scholar 

  22. Sawa, H., Kouike, H. & Okano, H. Components of the SWI/SNF complex are required for asymmetric cell division in C. elegans. Mol. Cell 6, 617–624 (2000).

    Article  CAS  PubMed  Google Scholar 

  23. Shibata, Y., Uchida, M., Takeshita, H., Nishiwaki, K. & Sawa, H. Multiple functions of PBRM-1/Polybromo- and LET-526/Osa-containing chromatin remodeling complexes in C. elegans development. Dev. Biol. 361, 349–357 (2012).

    Article  CAS  PubMed  Google Scholar 

  24. Yoo, A. S., Staahl, B. T., Chen, L. & Crabtree, G. R. MicroRNA-mediated switching of chromatin-remodelling complexes in neural development. Nature 460, 642–646 (2009). This paper defines the regulatory circuitry that underlies the switch of BAF subunits as neural progenitors exit mitosis and differentiate into neurons. Later studies by this group found that recapitulation of this switch would change human fibroblasts into neurons.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Yoo, A. S. et al. MicroRNA-mediated conversion of human fibroblasts to neurons. Nature 476, 228–231 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Tea, J. S. & Luo, L. The chromatin remodeling factor Bap55 functions through the TIP60 complex to regulate olfactory projection neuron dendrite targeting. Neural Dev. 6, 5 (2011). The authors conducted a daring screen for mutations that would result in the perfect retargeting of the olfactory neuron dendritic tree to a different glomerulus. Surprisingly, the only gene in which a mutation produces perfect retargeting is the homologue of BAF53 , which switches during the course of mammalian neural development from BAF53A to BAF53B . These studies suggest that the pathogenic mechanisms involved in diseases associated with mutations in BAF complexes might be due to defects in dendritic targeting.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Parrish, J. Z., Kim, M. D., Jan, L. Y. & Jan, Y. N. Genome-wide analyses identify transcription factors required for proper morphogenesis of Drosophila sensory neuron dendrites. Genes Dev. 20, 820–835 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Santen, G. W. et al. Mutations in SWI/SNF chromatin remodeling complex gene ARID1B cause Coffin–Siris syndrome. Nature Genet. 44, 379–380 (2012).

    Article  CAS  PubMed  Google Scholar 

  29. Tsurusaki, Y. et al. Mutations affecting components of the SWI/SNF complex cause Coffin–Siris syndrome. Nature Genet. 44, 376–378 (2012).

    Article  CAS  PubMed  Google Scholar 

  30. Van Houdt, J. K. et al. Heterozygous missense mutations in SMARCA2 cause Nicolaides–Baraitser syndrome. Nature Genet. 44, 445–449 (2012).

    Article  CAS  PubMed  Google Scholar 

  31. Wolff, D. et al. In-frame deletion and missense mutations of the C-terminal helicase domain of SMARCA2 in three patients with Nicolaides–Baraitser syndrome. Mol. Syndromol. 2, 237–244 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Halgren, C. et al. Corpus callosum abnormalities, intellectual disability, speech impairment, and autism in patients with haploinsufficiency of ARID1B. Clin. Genet. 82, 248–255 (2011).

    Article  CAS  PubMed  Google Scholar 

  33. Backx, L., Seuntjens, E., Devriendt, K., Vermeesch, J. & Van Esch, H. A balanced translocation t(6;14)(q25.3;q13.2) leading to reciprocal fusion transcripts in a patient with intellectual disability and agenesis of corpus callosum. Cytogenet. Genome Res. 132, 135–143 (2011).

    Article  CAS  PubMed  Google Scholar 

  34. Hoyer, J. et al. Haploinsufficiency of ARID1B, a member of the SWI/SNF-A chromatin-remodeling complex, is a frequent cause of intellectual disability. Am. J. Hum. Genet. 90, 565–572 (3012).

    Article  CAS  Google Scholar 

  35. Neale, B. M. et al. Patterns and rates of exonic de novo mutations in autism spectrum disorders. Nature 485, 242–245 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. O'Roak, B. J. et al. Sporadic autism exomes reveal a highly interconnected protein network of de novo mutations. Nature 485, 246–250 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Zoghbi, H. Y. Postnatal neurodevelopmental disorders: meeting at the synapse? Science 302, 826–830 (2003).

    Article  CAS  PubMed  Google Scholar 

  38. Zoghbi, H. Y. & Bear, M. F. Synaptic dysfunction in neurodevelopmental disorders associated with autism and intellectual disabilities. Cold Spring Harb. Perspect. Biol. 18 Jan 2012 (10.1101/cshperspect.a009886).

  39. Koga, M. et al. Involvement of SMARCA2/BRM in the SWI/SNF chromatin-remodeling complex in schizophrenia. Hum. Mol. Genet. 18, 2483–2494 (2009).

    Article  CAS  PubMed  Google Scholar 

  40. Loe-Mie, Y. et al. SMARCA2 and other genome-wide supported schizophrenia-associated genes: regulation by REST/NRSF, network organization and primate-specific evolution. Hum. Mol. Genet. 19, 2841–2857 (2010).

    Article  CAS  PubMed  Google Scholar 

  41. Nishiyama, M. et al. CHD8 suppresses p53-mediated apoptosis through histone H1 recruitment during early embryogenesis. Nature Cell Biol. 11, 172–182 (2009).

    Article  CAS  PubMed  Google Scholar 

  42. Thompson, B. A., Tremblay, V., Lin, G. & Bochar, D. A. CHD8 is an ATP-dependent chromatin remodeling factor that regulates beta-catenin target genes. Mol. Cell. Biol. 28, 3894–3904 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Sakamoto, I. et al. A novel β-catenin-binding protein inhibits β-catenin-dependent TCF activation and axis formation. J. Biol. Chem. 275, 32871–32878 (2000).

    Article  CAS  PubMed  Google Scholar 

  44. Nishiyama, M., Skoultchi, A. I. & Nakayama, K. I. Histone H1 recruitment by CHD8 is essential for suppression of the Wnt-β-catenin signaling pathway. Mol. Cell. Biol. 32, 501–512 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Heasman, J., Kofron, M. & Wylie, C. β-catenin signaling activity dissected in the early Xenopus embryo: a novel antisense approach. Dev. Biol. 222, 124–134 (2000).

    Article  CAS  PubMed  Google Scholar 

  46. Chenn, A. & Walsh, C. A. Regulation of cerebral cortical size by control of cell cycle exit in neural precursors. Science 297, 365–369 (2002).

    Article  CAS  PubMed  Google Scholar 

  47. Brault, V. et al. Inactivation of the β-catenin gene by Wnt1-Cre-mediated deletion results in dramatic brain malformation and failure of craniofacial development. Development 128, 1253–1264 (2001).

    CAS  PubMed  Google Scholar 

  48. Ille, F. & Sommer, L. Wnt signaling: multiple functions in neural development. Cell. Mol. Life Sci. 62, 1100–1108 (2005).

    Article  CAS  PubMed  Google Scholar 

  49. Barker, N. et al. The chromatin remodelling factor Brg-1 interacts with β-catenin to promote target gene activation. EMBO J. 20, 4935–4943 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. De Ferrari, G. V. & Moon, R. T. The ups and downs of WNT signaling in prevalent neurological disorders. Oncogene 25, 7545–7553 (2006).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Talkowski, M. E. et al. Sequencing chromosomal abnormalities reveals neurodevelopmental loci that confer risk across diagnostic boundaries. Cell 149, 525–537 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Sanders, S. J. et al. De novo mutations revealed by whole-exome sequencing are strongly associated with autism. Nature 485, 237–241 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Williams, C. A., Dagli, A. & Battaglia, A. Genetic disorders associated with macrocephaly. Am. J. Med. Genet. A 146, 2023–2037 (2008).

    Article  CAS  Google Scholar 

  55. Henikoff, S. & Shilatifard, A. Histone modification: cause or cog? Trends Genet. 27, 389–396 (2011).

    Article  CAS  PubMed  Google Scholar 

  56. Taunton, J., Hassig, C. A. & Schreiber, S. L. A mammalian histone deacetylase related to the yeast transcriptional regulator Rpd3p. Science 272, 408–411 (1996). The authors describe the first evidence that epigenetic modifications are reversible by enzymes that remove or erase the modification. This paper describes the discovery of HDAC and initiated the idea that epigenetic modification was a dynamic state rather than an irreversible process.

    Article  CAS  PubMed  Google Scholar 

  57. Shi, Y. et al. Histone demethylation mediated by the nuclear amine oxidase homolog LSD1. Cell 119, 941–953 (2004).

    Article  CAS  PubMed  Google Scholar 

  58. Wang, Z. et al. Genome-wide mapping of HATs and HDACs reveals distinct functions in active and inactive genes. Cell 138, 1019–1031 (2009). The authors use the chromatin immunoprecipitation followed by high-throughput sequencing (ChIP–seq) method, which they invented, to demonstrate that histone acetylases and deacetylases bind to the same sites over the genome indicating that these modifications are in dynamic equilibrium and hence subject to regulatory mechanisms.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Hathaway, N. A. et al. Dynamics and memory of heterochromatin in living cells. Cell 149, 1447–1460 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Egger, G., Liang, G., Aparicio, A. & Jones, P. A. Epigenetics in human disease and prospects for epigenetic therapy. Nature 429, 457–463 (2004).

    Article  CAS  PubMed  Google Scholar 

  61. Cross, N. C. Histone modification defects in developmental disorders and cancer. Oncotarget 3, 3–4 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  62. Kramer, J. M. & van Bokhoven, H. Genetic and epigenetic defects in mental retardation. Int. J. Biochem. Cell Biol. 41, 96–107 (2009).

    Article  CAS  PubMed  Google Scholar 

  63. Margueron, R. et al. Role of the Polycomb protein EED in the propagation of repressive histone marks. Nature 461, 762–767 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Hansen, K. H. et al. A model for transmission of the H3K27me3 epigenetic mark. Nature Cell Biol. 10, 1291–1300 (2008).

    Article  CAS  PubMed  Google Scholar 

  65. Francis, N. J., Kingston, R. E. & Woodcock, C. L. Chromatin compaction by a Polycomb group protein complex. Science 306, 1574–1577 (2004).

    Article  CAS  PubMed  Google Scholar 

  66. O'Carroll, D. et al. The Polycomb-group gene Ezh2 is required for early mouse development. Mol. Cell. Biol. 21, 4330–4336 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Boyer, L. A. et al. Polycomb complexes repress developmental regulators in murine embryonic stem cells. Nature 441, 349–353 (2006).

    Article  CAS  PubMed  Google Scholar 

  68. Chamberlain, S. J., Yee, D. & Magnuson, T. Polycomb repressive complex 2 is dispensable for maintenance of embryonic stem cell pluripotency. Stem Cells 26, 1496–1505 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Pereira, J. D. et al. Ezh2, the histone methyltransferase of PRC2, regulates the balance between self-renewal and differentiation in the cerebral cortex. Proc. Natl Acad. Sci. USA 107, 15957–15962 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  70. Hirabayashi, Y. et al. Polycomb limits the neurogenic competence of neural precursor cells to promote astrogenic fate transition. Neuron 63, 600–613 (2009).

    Article  CAS  PubMed  Google Scholar 

  71. Li, X., Han, Y. & Xi, R. Polycomb group genes Psc and Su(z)2 restrict follicle stem cell self-renewal and extrusion by controlling canonical and noncanonical Wnt signaling. Genes Dev. 24, 933–946 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Ho, L. et al. esBAF facilitates pluripotency by conditioning the genome for LIF/STAT3 signalling and by regulating polycomb function. Nature Cell Biol. 13, 903–913 (2011).

    Article  CAS  PubMed  Google Scholar 

  73. Cohen, M. M. Jr. Mental deficiency, alterations in performance, and CNS abnormalities in overgrowth syndromes. Am. J. Med. Genet. C 117, 49–56 (2003).

    Article  Google Scholar 

  74. Gibson, W. T. et al. Mutations in EZH2 cause Weaver syndrome. Am. J. Hum. Genet. 90, 110–118 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Tatton-Brown, K. et al. Germline mutations in the oncogene EZH2 cause Weaver syndrome and increased human height. Oncotarget 2, 1127–1133 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  76. Wang, A. H. & Yang, X. J. Histone deacetylase 4 possesses intrinsic nuclear import and export signals. Mol. Cell. Biol. 21, 5992–6005 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Bolger, T. A. & Yao, T. P. Intracellular trafficking of histone deacetylase 4 regulates neuronal cell death. J. Neurosci. 25, 9544–9553 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Lahm, A. et al. Unraveling the hidden catalytic activity of vertebrate class IIa histone deacetylases. Proc. Natl Acad. Sci. USA 104, 17335–17340 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  79. Chan, J. K., Sun, L., Yang, X. J., Zhu, G. & Wu, Z. Functional characterization of an amino-terminal region of HDAC4 that possesses MEF2 binding and transcriptional repressive activity. J. Biol. Chem. 278, 23515–23521 (2003).

    Article  CAS  PubMed  Google Scholar 

  80. Zhang, C. L., McKinsey, T. A. & Olson, E. N. Association of class II histone deacetylases with heterochromatin protein 1: potential role for histone methylation in control of muscle differentiation. Mol. Cell. Biol. 22, 7302–7312 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Kim, M. S. et al. An essential role for histone deacetylase 4 in synaptic plasticity and memory formation. J. Neurosci. 32, 10879–10886 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Vecsey, C. G. et al. Histone deacetylase inhibitors enhance memory and synaptic plasticity via CREB:CBP-dependent transcriptional activation. J. Neurosci. 27, 6128–6140 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Sando, R. et al. HDAC4 governs a transcriptional program essential for synaptic plasticity and memory. Cell 151, 821–834 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Shalizi, A. K. & Bonni, A. Brawn for brains: the role of MEF2 proteins in the developing nervous system. Curr. Top. Dev. Biol. 69, 239–266 (2005).

    Article  CAS  PubMed  Google Scholar 

  86. Flavell, S. W. et al. Genome-wide analysis of MEF2 transcriptional program reveals synaptic target genes and neuronal activity-dependent polyadenylation site selection. Neuron 60, 1022–1038 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Williams, S. R. et al. Haploinsufficiency of HDAC4 causes brachydactyly mental retardation syndrome, with brachydactyly type E, developmental delays, and behavioral problems. Am. J. Hum. Genet. 87, 219–228 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Morris, B. et al. Dose dependent expression of HDAC4 causes variable expressivity in a novel inherited case of brachydactyly mental retardation syndrome. Am. J. Med. Genet. A 158, 2015–2020 (2012).

    Article  CAS  Google Scholar 

  89. Ramocki, M. B. & Zoghbi, H. Y. Failure of neuronal homeostasis results in common neuropsychiatric phenotypes. Nature 455, 912–918 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Korzus, E., Rosenfeld, M. G. & Mayford, M. CBP histone acetyltransferase activity is a critical component of memory consolidation. Neuron 42, 961–972 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Levenson, J. M. & Sweatt, J. D. Epigenetic mechanisms in memory formation. Nature Rev. Neurosci. 6, 108–118 (2005).

    Article  CAS  Google Scholar 

  92. Guan, Z. et al. Integration of long-term-memory-related synaptic plasticity involves bidirectional regulation of gene expression and chromatin structure. Cell 111, 483–493 (2002).

    Article  CAS  PubMed  Google Scholar 

  93. Gupta, S. et al. Histone methylation regulates memory formation. J. Neurosci. 30, 3589–3599 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Holtmaat, A. & Svoboda, K. Experience-dependent structural synaptic plasticity in the mammalian brain. Nature Rev. Neurosci. 10, 647–658 (2009).

    Article  CAS  Google Scholar 

  95. Tamkun, J. W. et al. Brahma: a regulator of Drosophila homeotic genes structurally related to the yeast transcriptional activator SNF2/SWI2. Cell 68, 561–572 (1992). The authors define the opposition between Polycomb and BAF or BAP complexes, which is likely to be a crucial underlying mechanism of both human malignancy and neurodevelopmental disorders.

    Article  CAS  PubMed  Google Scholar 

  96. Wilson, B. G. et al. Epigenetic antagonism between polycomb and SWI/SNF complexes during oncogenic transformation. Cancer Cell 18, 316–328 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Kim, J. E. et al. Investigating synapse formation and function using human pluripotent stem cell-derived neurons. Proc. Natl Acad. Sci. USA 108, 3005–3010 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  98. Brennand, K. J., Simone, A., Tran, N. & Gage, F. H. Modeling psychiatric disorders at the cellular and network levels. Mol. Psychiatry 17, 1239–1253 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Brennand, K. J. et al. Modelling schizophrenia using human induced pluripotent stem cells. Nature 473, 221–225 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Kriks, S. et al. Dopamine neurons derived from human ES cells efficiently engraft in animal models of Parkinson's disease. Nature 480, 547–551 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Caiazzo, M. et al. Direct generation of functional dopaminergic neurons from mouse and human fibroblasts. Nature 476, 224–227 (2011).

    Article  CAS  PubMed  Google Scholar 

  102. Kadoch, C., Hargreaves, D. Hodges, C. & Crabtree, G. R. Nature Genet. (in the press).

  103. Shane, A. H. & Pollack, J. R. The spectrum of SWI/SNF mutations, ubiquitous in human cancers. PLoS ONE 8, e55119 (2013).

    Article  CAS  Google Scholar 

  104. Wiegand, K. C. et al. ARID1A mutations in endometriosis-associated ovarian carcinomas. N. Engl. J. Med. 363, 1532–1543 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Wu, J. N. & Roberts, C. W. ARID1A mutations in cancer: another epigenetic tumor suppressor? Cancer Discov. 3, 35–43 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Shih, A. H., Abdel-Wahab, O., Patel, J. P. & Levine, R. L. The role of mutations in epigenetic regulators in myeloid malignancies. Nature Rev. Cancer 12, 599–612 (2012).

    Article  CAS  Google Scholar 

  107. Gied, J. Brain development IX: human brain growth. Am. J. Psychiatry 156, 4 (1999).

    Article  Google Scholar 

  108. Kleefstra, T. et al. Disruption of an EHMT1-associated chromatin-modification module causes intellectual disability. Am. J. Hum. Genet. 91, 73–82 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Vissers, L. E. et al. Mutations in a new member of the chromodomain gene family cause CHARGE syndrome. Nature Genet. 36, 955–957 (2004).

    Article  CAS  PubMed  Google Scholar 

  110. Iwase, S. et al. ATRX ADD domain links an atypical histone methylation recognition mechanism to human mental-retardation syndrome. Nature Struct. Mol. Biol. 18, 769–776 (2011).

    Article  CAS  Google Scholar 

  111. Roelfsema, J. H. et al. Genetic heterogeneity in Rubinstein–Taybi syndrome: mutations in both the CBP and EP300 genes cause disease. Am. J. Hum. Genet. 76, 572–580 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Petrij, F. et al. Rubinstein–Taybi syndrome caused by mutations in the transcriptional co-activator CBP. Nature 376, 348–351 (1995).

    Article  CAS  PubMed  Google Scholar 

  113. Clayton-Smith, J. et al. Whole-exome-sequencing identifies mutations in histone acetyltransferase gene KAT6B in individuals with the Say–Barber–Biesecker variant of Ohdo syndrome. Am. J. Hum. Genet. 89, 675–681 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Kleefstra, T. et al. Loss-of-function mutations in euchromatin histone methyl transferase 1 (EHMT1) cause the 9q34 subtelomeric deletion syndrome. Am. J. Hum. Genet. 79, 370–377 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Jones, W. D. et al. De novo mutations in MLL cause Wiedemann–Steiner syndrome. Am. J. Hum. Genet. 91, 358–364 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Hannibal, M. C. et al. Spectrum of MLL2 (ALR) mutations in 110 cases of Kabuki syndrome. Am. J. Med Genet. A 155, 1511–1516 (2011).

    Article  CAS  PubMed Central  Google Scholar 

  117. Ng, S. B. et al. Exome sequencing identifies MLL2 mutations as a cause of Kabuki syndrome. Nature Genet. 42, 790–793 (2010).

    Article  CAS  PubMed  Google Scholar 

  118. Jensen, L. R. et al. Mutations in the JARID1C gene, which is involved in transcriptional regulation and chromatin remodeling, cause X-linked mental retardation. Am. J. Hum. Genet. 76, 227–236 (2005).

    Article  CAS  PubMed  Google Scholar 

  119. Najmabadi, H. et al. Deep sequencing reveals 50 novel genes for recessive cognitive disorders. Nature 478, 57–63 (2011).

    Article  CAS  PubMed  Google Scholar 

  120. Koivisto, A. M. et al. Screening of mutations in the PHF8 gene and identification of a novel mutation in a Finnish family with XLMR and cleft lip/cleft palate. Clin. Genet. 72, 145–149 (2007).

    Article  CAS  PubMed  Google Scholar 

  121. Laumonnier, F. et al. Mutations in PHF8 are associated with X linked mental retardation and cleft lip/cleft palate. J. Med. Genet. 42, 780–786 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Froyen, G. et al. Submicroscopic duplications of the hydroxysteroid dehydrogenase HSD17B10 and the E3 ubiquitin ligase HUWE1 are associated with mental retardation. Am. J. Hum. Genet. 82, 432–443 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Amir, R. E. et al. Rett syndrome is caused by mutations in X-linked MECP2, encoding methyl-CpG-binding protein 2. Nature Genet. 23, 185–188 (1999).

    Article  CAS  PubMed  Google Scholar 

  124. Moretti, P. & Zoghbi, H. Y. MeCP2 dysfunction in Rett syndrome and related disorders. Curr. Opin. Genet. Dev. 16, 276–281 (2006).

    Article  CAS  PubMed  Google Scholar 

  125. Risheg, H. et al. A recurrent mutation in MED12 leading to R961W causes Opitz–Kaveggia syndrome. Nature Genet. 39, 451–453 (2007).

    Article  CAS  PubMed  Google Scholar 

  126. Schwartz, C. E. et al. The original Lujan syndrome family has a novel missense mutation (p.N1007S) in the MED12 gene. J. Med. Genet. 44, 472–477 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Hashimoto, S. et al. MED23 mutation links intellectual disability to dysregulation of immediate early gene expression. Science 333, 1161–1163 (2011).

    Article  CAS  PubMed  Google Scholar 

  128. Kim, H. G. et al. Translocations disrupting PHF21A in the Potocki–Shaffer-syndrome region are associated with intellectual disability and craniofacial anomalies. Am. J. Hum. Genet. 91, 56–72 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We apologize to colleagues whose work we did not cite owing to lack of space or unintentional oversight. This work was supported by the Howard Hughes Medical Foundation, a grant from the California Institute for Regenerative Medicine and US National Institutes of Health grants NS046789 and CA163915 to G.R.C.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gerald R. Crabtree.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

FURTHER INFORMATION

Gerald R. Crabtree's homepage

PowerPoint slides

Supplementary information

Supplementary Table 1

Chromatin regulators mutated in mental disorders (PDF 277 kb)

Supplementary Information

Tables of mutations from Box 1 (XLS 100 kb)

Glossary

Exome sequencing

Targeted sequencing of protein-coding regions of the genome (that is, exons). It is a cheaper yet still effective alternative to whole-genome sequencing to identify clinically relevant gene variations that are responsible for both Mendelian and common diseases but does not detect mutations in non-coding regions of the genome

Glomerulus

The glomerulus is a spherical structure located in the olfactory bulb. Each glomerulus is thought to receive input from olfactory receptor neurons (expressing only one type of olfactory receptor). They relay this information into higher brain structure through projection neurons.

Autosomal dominant disease

Refers to a disease arising from mutations on non-sex chromosomes that are genetically dominant.

Microcephaly

Microcephaly is a significantly smaller head measurement. It can be caused by an abnormal brain size due to loss of any number of cell types or brain features and can even be caused by abnormal ventricular spaces or cerebral fluid.

Haploinsufficiency

When the product of one normal allele of a gene is not sufficient to allow the normal function of a gene to be executed. This is another possible cause of genetic dominant diseases.

Frameshift indel

An insertion and/or deletion mutation that changes the reading frame of a protein and creates an altered gene product.

Dominant-negative

Interference with the function of the normal allele of a gene by a mutated allele. This usually occurs when the mutant product can still interact with the same elements as the wild-type product, but block some aspect of its function. This is another possible cause of genetic dominant diseases.

Corpus callosum

A wide flat bundle of neural fibres that connects the left and right cerebral hemispheres and facilitates interhemispheric communication.

De novo mutations

Alterations in genes that are present for the first time in one family member as a result of a mutation in a germ cell (that is, an egg or sperm) of one of the parents or in the fertilized egg itself.

WNT/β-catenin signalling

The binding of WNT ligand to receptor Frizzled leads to a cascade of events allowing for β-catenin (an integral cell–cell adhesion adaptor protein as well as transcriptional co-regulator) stabilization, nuclear translocation and transcriptional activation. The WNT/β-catenin pathway integrates signals from many other pathways, including retinoic acid, FGF, TGFβ and BMP in many different cell types and tissues.

Single-nucleotide polymorphisms

(SNPs). DNA sequence variations that occur when a single nucleotide (A, T, C or G) in the genome or other shared sequence differs between members of a biological species or paired chromosomes in an individual.

Linker histone

The linker histone H1 binds the nucleosome and the entry or exit sites of the DNA, allowing for the formation of higher-order chromatin structures that are thought to lead to chromatin compaction and gene repression.

Macrocephaly

Macrocephaly is a significantly larger than average head circumference measurement. It can be caused by an abnormal brain size due to gain of any number of cell types or brain features and can even be caused by abnormal ventricular spaces or cerebral fluid.

Polycomb group proteins

These conserved proteins form multimeric complexes that exert their functions by modifying chromatin structure and by regulating the deposition and recognition of multiple post-translational histone modifications. Their epigenetic role appears to arise from their ability to propagate a repressive chromatin modification over several kilobases of DNA.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Ronan, J., Wu, W. & Crabtree, G. From neural development to cognition: unexpected roles for chromatin. Nat Rev Genet 14, 347–359 (2013). https://doi.org/10.1038/nrg3413

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrg3413

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research