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:

Integrative regulation of physiology by histone deacetylase 3

Abstract

Cell-type-specific gene expression is physiologically modulated by the binding of transcription factors to genomic enhancer sequences, to which chromatin modifiers such as histone deacetylases (HDACs) are recruited. Drugs that inhibit HDACs are in clinical use but lack specificity. HDAC3 is a stoichiometric component of nuclear receptor co-repressor complexes whose enzymatic activity depends on this interaction. HDAC3 is required for many aspects of mammalian development and physiology, for example, for controlling metabolism and circadian rhythms. In this Review, we discuss the mechanisms by which HDAC3 regulates cell type-specific enhancers, the structure of HDAC3 and its function as part of nuclear receptor co-repressors, its enzymatic activity and its post-translational modifications. We then discuss the plethora of tissue-specific physiological functions of HDAC3.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: HDAC3 is a core component of nuclear receptor co-repressor complexes that modulate nuclear receptor-mediated transcription.
Fig. 2: HDAC3 suppresses liver metabolism and circadian clock genes through distinct enhancer complexes.
Fig. 3: HDAC3 primes thermogenic gene transcription in brown adipose tissue.
Fig. 4: HDAC3 influences cardiac development through deacetylase-independent mechanisms.
Fig. 5: HDAC3 controls brain development, glial cell fate and the formation of long-term memory.
Fig. 6: HDAC3 regulates lung development, intestinal homeostasis, pancreatic β-cell insulin secretion and skeletal muscle metabolism.
Fig. 7: HDAC3 regulates distinct mouse tissue-specific gene expression programmes.

Similar content being viewed by others

References

  1. Ong, C.-T. & Corces, V. G. Enhancer function: new insights into the regulation of tissue-specific gene expression. Nat. Rev. Genet. 12, 283–293 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Spitz, F. & Furlong, E. E. M. Transcription factors: from enhancer binding to developmental control. Nat. Rev. Genet. 13, 613–626 (2012).

    CAS  PubMed  Google Scholar 

  3. Glass, C. K. & Natoli, G. Molecular control of activation and priming in macrophages. Nat. Immunol. 17, 26–33 (2016).

    CAS  PubMed  Google Scholar 

  4. Carroll, J. S. et al. Chromosome-wide mapping of estrogen receptor binding reveals long-range regulation requiring the forkhead protein FoxA1. Cell 122, 33–43 (2005).

    CAS  PubMed  Google Scholar 

  5. Lefterova, M. I. et al. PPARγ and C/EBP factors orchestrate adipocyte biology via adjacent binding on a genome-wide scale. Genes Dev. 22, 2941–2952 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Biddie, S. C. et al. Transcription factor AP1 potentiates chromatin accessibility and glucocorticoid receptor binding. Mol. Cell 43, 145–155 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. John, S. et al. Chromatin accessibility pre-determines glucocorticoid receptor binding patterns. Nat. Genet. 43, 264–268 (2011).

    CAS  PubMed  Google Scholar 

  8. Grøntved, L. et al. C/EBP maintains chromatin accessibility in liver and facilitates glucocorticoid receptor recruitment to steroid response elements. EMBO J. 32, 1568–1583 (2013).

    PubMed  PubMed Central  Google Scholar 

  9. Madsen, M. S., Siersbæk, R., Boergesen, M., Nielsen, R. & Mandrup, S. Peroxisome proliferator-activated receptor γ and C/EBPα synergistically activate key metabolic adipocyte genes by assisted loading. Mol. Cell. Biol. 34, 939–954 (2014).

    PubMed  PubMed Central  Google Scholar 

  10. Soccio, R. E. et al. Genetic variation determines PPARγ function and anti-diabetic drug response in vivo. Cell 162, 33–44 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Yang, X.-J. & Seto, E. Lysine acetylation: codified crosstalk with other posttranslational modifications. Mol. Cell 31, 449–461 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Verdin, E. & Ott, M. 50 years of protein acetylation: from gene regulation to epigenetics, metabolism and beyond. Nat. Rev. Mol. Cell Biol. 16, 258–264 (2015). This is an excellent review of the role of protein acetylation in biology.

    CAS  PubMed  Google Scholar 

  13. Bannister, A. J. & Kouzarides, T. The CBP co-activator is a histone acetyltransferase. Nature 384, 641–643 (1996).

    CAS  PubMed  Google Scholar 

  14. Ogryzko, V. V., Schiltz, R. L., Russanova, V., Howard, B. H. & Nakatani, Y. The transcriptional coactivators p300 and CBP are histone acetyltransferases. Cell 87, 953–959 (1996).

    CAS  PubMed  Google Scholar 

  15. Yang, X. J., Ogryzko, V. V., Nishikawa, J., Howard, B. H. & Nakatani, Y. A p300/CBP-associated factor that competes with the adenoviral oncoprotein E1A. Nature 382, 319–324 (1996).

    CAS  PubMed  Google Scholar 

  16. Spencer, T. E. et al. Steroid receptor coactivator-1 is a histone acetyltransferase. Nature 389, 194–198 (1997).

    CAS  PubMed  Google Scholar 

  17. Haberland, M., Montgomery, R. L. & Olson, E. N. The many roles of histone deacetylases in development and physiology: implications for disease and therapy. Nat. Rev. Genet. 10, 32–42 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Taunton, J., Hassig, C. A. & Schreiber, S. L. A mammalian histone deacetylase related to the yeast transcriptional regulator Rpd3p. Sci. 272, 408–411 (1996).

    CAS  Google Scholar 

  19. Yang, X.-J. & Seto, E. The Rpd3/Hda1 family of lysine deacetylases: from bacteria and yeast to mice and men. Nat. Rev. Mol. Cell Biol. 9, 206–218 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Kadosh, D. & Struhl, K. Histone deacetylase activity of Rpd3 is important for transcriptional repression in vivo. Genes Dev. 12, 797–805 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Rundlett, S. E. et al. HDA1 and RPD3 are members of distinct yeast histone deacetylase complexes that regulate silencing and transcription. Proc. Natl Acad. Sci. USA 93, 14503–14508 (1996).

    CAS  PubMed  Google Scholar 

  22. Gregoretti, I. V., Lee, Y. M. & Goodson, H. V. Molecular evolution of the histone deacetylase family: functional implications of phylogenetic analysis. J. Mol. Biol. 338, 17–31 (2004).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  24. Schwer, B. & Verdin, E. Conserved metabolic regulatory functions of sirtuins. Cell. Metab. 7, 104–112 (2008).

    CAS  PubMed  Google Scholar 

  25. Chang, H.-C. & Guarente, L. SIRT1 and other sirtuins in metabolism. Trends Endocrinol. Metab. 25, 138–145 (2014).

    CAS  PubMed  Google Scholar 

  26. Chalkiadaki, A. & Guarente, L. Sirtuins mediate mammalian metabolic responses to nutrient availability. Nat. Rev. Endocrinol. 8, 287–296 (2012).

    CAS  PubMed  Google Scholar 

  27. Seto, E. & Yoshida, M. Erasers of histone acetylation: the histone deacetylase enzymes. Cold Spring Harb. Perspect. Biol. 6, a018713 (2014).

    PubMed  PubMed Central  Google Scholar 

  28. West, A. C. & Johnstone, R. W. New and emerging HDAC inhibitors for cancer treatment. J. Clin. Invest. 124, 30–39 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Li, J. et al. Both corepressor proteins SMRT and N-CoR exist in large protein complexes containing HDAC3. EMBO J. 19, 4342–4350 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Wen, Y. D. et al. The histone deacetylase-3 complex contains nuclear receptor corepressors. Proc. Natl Acad. Sci. USA 97, 7202–7207 (2000).

    CAS  PubMed  Google Scholar 

  31. Guenther, M. G. et al. A core SMRT corepressor complex containing HDAC3 and TBL1, a WD40-repeat protein linked to deafness. Genes Dev. 14, 1048–1057 (2000). References 29–31 identify the stoichiometric relationship between nuclear receptor co-repressors and HDAC3.

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Everett, L. J. & Lazar, M. A. Cell-specific integration of nuclear receptor function at the genome. Wiley Interdiscip. Rev. Syst. Biol. Med. 5, 615–629 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Francis, G. A., Fayard, E., Picard, F. & Auwerx, J. Nuclear receptors and the control of metabolism. Annu. Rev. Physiol. 65, 261–311 (2003).

    CAS  PubMed  Google Scholar 

  34. Perissi, V. & Rosenfeld, M. G. Controlling nuclear receptors: the circular logic of cofactor cycles. Nat. Rev. Mol. Cell Biol. 6, 542–554 (2005).

    CAS  PubMed  Google Scholar 

  35. Hong, S. H. et al. Nuclear receptors and metabolism: from feast to famine. Diabetologia 57, 860–867 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Glass, C. K. & Rosenfeld, M. G. The coregulator exchange in transcriptional functions of nuclear receptors. Genes Dev. 14, 121–141 (2000).

    CAS  PubMed  Google Scholar 

  37. McKenna, N. J., Lanz, R. B. & O’Malley, B. W. Nuclear receptor coregulators: cellular and molecular biology. Endocr. Rev. 20, 321–344 (1999).

    CAS  PubMed  Google Scholar 

  38. Hu, X. & Lazar, M. A. The CoRNR motif controls the recruitment of corepressors by nuclear hormone receptors. Nature 402, 93–96 (1999).

    CAS  PubMed  Google Scholar 

  39. Perissi, V., Jepsen, K., Glass, C. K. & Rosenfeld, M. G. Deconstructing repression: evolving models of co-repressor action. Nat. Rev. Genet. 11, 109–123 (2010).

    CAS  PubMed  Google Scholar 

  40. Yoon, H. G. et al. Purification and functional characterization of the human N-CoR complex: the roles of HDAC3, TBL1 and TBLR1. EMBO J. 22, 1336–1346 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Zhang, J., Kalkum, M., Chait, B. T. & Roeder, R. G. The N-CoR-HDAC3 nuclear receptor corepressor complex inhibits the JNK pathway through the integral subunit GPS2. Mol. Cell 9, 611–623 (2002).

    CAS  PubMed  Google Scholar 

  42. Kao, H. Y., Downes, M., Ordentlich, P. & Evans, R. M. Isolation of a novel histone deacetylase reveals that class I and class II deacetylases promote SMRT-mediated repression. Genes Dev. 14, 55–66 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Heinzel, T. et al. A complex containing N-CoR, mSin3 and histone deacetylase mediates transcriptional repression. Nature 387, 43–48 (1997).

    CAS  PubMed  Google Scholar 

  44. Nagy, L. et al. Nuclear receptor repression mediated by a complex containing SMRT, mSin3A, and histone deacetylase. Cell 89, 373–380 (1997).

    CAS  PubMed  Google Scholar 

  45. Huang, E. Y. et al. Nuclear receptor corepressors partner with class II histone deacetylases in a Sin3-independent repression pathway. Genes Dev. 14, 45–54 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Perissi, V. et al. Molecular determinants of nuclear receptor-corepressor interaction. Genes Dev. 13, 3198–3208 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Guenther, M. G., Barak, O. & Lazar, M. A. The SMRT and N-CoR corepressors are activating cofactors for histone deacetylase 3. Mol. Cell. Biol. 21, 6091–6101 (2001). This paper reports that HDAC3 activity requires nuclear receptor co-repressors.

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Ishizuka, T. & Lazar, M. A. The nuclear receptor corepressor deacetylase activating domain is essential for repression by thyroid hormone receptor. Mol. Endocrinol. 19, 1443–1451 (2005).

    CAS  PubMed  Google Scholar 

  49. Codina, A. et al. Structural insights into the interaction and activation of histone deacetylase 3 by nuclear receptor corepressors. Proc. Natl Acad. Sci. USA 102, 6009–6014 (2005).

    CAS  PubMed  Google Scholar 

  50. Aasland, R., Stewart, A. F. & Gibson, T. The SANT domain: a putative DNA-binding domain in the SWI-SNF and ADA complexes, the transcriptional co-repressor N-CoR and TFIIIB. Trends Biochem. Sci. 21, 87–88 (1996).

    CAS  PubMed  Google Scholar 

  51. Boyer, L. A. et al. Essential role for the SANT domain in the functioning of multiple chromatin remodeling enzymes. Mol. Cell 10, 935–942 (2002).

    CAS  PubMed  Google Scholar 

  52. Boyer, L. A., Latek, R. R. & Peterson, C. L. The SANT domain: a unique histone-tail-binding module? Nat. Rev. Mol. Cell Biol. 5, 158–163 (2004).

    CAS  PubMed  Google Scholar 

  53. Yu, J., Li, Y., Ishizuka, T., Guenther, M. G. & Lazar, M. A. A. SANT motif in the SMRT corepressor interprets the histone code and promotes histone deacetylation. EMBO J. 22, 3403–3410 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Watson, P. J., Fairall, L., Santos, G. M. & Schwabe, J. W. R. Structure of HDAC3 bound to co-repressor and inositol tetraphosphate. Nature 481, 335–340 (2012). This paper reports the crystal structure of HDAC3 bound to a nuclear receptor DAD, identifying crucial interactions between the two proteins and a role for IP4 in maintaining them.

    CAS  PubMed  PubMed Central  Google Scholar 

  55. You, S.-H. et al. Nuclear receptor co-repressors are required for the histone-deacetylase activity of HDAC3 in vivo. Nat. Struct. Mol. Biol. 20, 182–187 (2013). This paper demonstrates that the enzymatic activity of HDAC3 requires nuclear receptor co-repressors in vivo.

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Adrain, C. & Freeman, M. New lives for old: evolution of pseudoenzyme function illustrated by iRhoms. Nat. Rev. Mol. Cell Biol. 13, 489–498 (2012).

    CAS  PubMed  Google Scholar 

  57. Guo, C. et al. Regulated clearance of histone deacetylase 3 protects independent formation of nuclear receptor corepressor complexes. J. Biol. Chem. 287, 12111–12120 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Guenther, M. G., Yu, J., Kao, G. D., Yen, T. J. & Lazar, M. A. Assembly of the SMRT-histone deacetylase 3 repression complex requires the TCP-1 ring complex. Genes Dev. 16, 3130–3135 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Sun, Z. et al. Deacetylase-independent function of HDAC3 in transcription and metabolism requires nuclear receptor corepressor. Mol. Cell 52, 769–782 (2013). This paper describes non-enzymatic functions of HDAC3.

    CAS  PubMed  Google Scholar 

  60. 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 

  61. 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 

  62. Alenghat, T. et al. Nuclear receptor corepressor and histone deacetylase 3 govern circadian metabolic physiology. Nature 456, 997–1000 (2008). This paper is the first to demonstrate a physiological role of the NCoR1–HDAC3 complex.

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Sengupta, N. & Seto, E. Regulation of histone deacetylase activities. J. Cell. Biochem. 93, 57–67 (2004).

    CAS  PubMed  Google Scholar 

  64. Lonard, D. M. & O’Malley, B. W. Nuclear receptor coregulators: judges, juries, and executioners of cellular regulation. Mol. Cell 27, 691–700 (2007).

    CAS  PubMed  Google Scholar 

  65. Tsai, S.-C. & Seto, E. Regulation of histone deacetylase 2 by protein kinase CK2. J. Biol. Chem. 277, 31826–31833 (2002).

    CAS  PubMed  Google Scholar 

  66. Zhang, X. et al. Histone deacetylase 3 (HDAC3) activity is regulated by interaction with protein serine/threonine phosphatase 4. Genes Dev. 19, 827–839 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Knutson, S. K. et al. Liver-specific deletion of histone deacetylase 3 disrupts metabolic transcriptional networks. EMBO J. 27, 1017–1028 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Villagra, A. et al. Histone deacetylase 3 down-regulates cholesterol synthesis through repression of lanosterol synthase gene expression. J. Biol. Chem. 282, 35457–35470 (2007).

    CAS  PubMed  Google Scholar 

  69. Sun, Z. et al. Hepatic Hdac3 promotes gluconeogenesis by repressing lipid synthesis and sequestration. Nat. Med. 18, 934–942 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. 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 

  71. Zamir, I., Zhang, J. & Lazar, M. A. Stoichiometric and steric principles governing repression by nuclear hormone receptors. Genes Dev. 11, 835–846 (1997).

    CAS  PubMed  Google Scholar 

  72. Hu, X., Li, Y. & Lazar, M. Determinants of CoRNR-dependent repression complex assembly on nuclear hormone receptors. Mol. Cell. Biol. 21, 1747–1758 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Bugge, A. et al. Rev-erbα and Rev-erbβ coordinately protect the circadian clock and normal metabolic function. Genes Dev. 26, 657–667 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Kim, Y. H. et al. Rev-erbα dynamically modulates chromatin looping to control circadian gene transcription. Science 359, 1274–1277 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Zhang, Y. et al. Gene regulation. Discrete functions of nuclear receptor Rev-erbα couple metabolism to the clock. Science 348, 1488–1492 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  76. Armour, S. M. et al. An HDAC3-PROX1 corepressor module acts on HNF4α to control hepatic triglycerides. Nat. Commun. 8, 549 (2017).

    PubMed  PubMed Central  Google Scholar 

  77. Emmett, M. J. et al. Histone deacetylase 3 prepares brown adipose tissue for acute thermogenic challenge. Nature 546, 544–548 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Enerbäck, S. et al. Mice lacking mitochondrial uncoupling protein are cold-sensitive but not obese. Nature 387, 90–94 (1997).

    PubMed  Google Scholar 

  79. Villena, J. A. et al. Orphan nuclear receptor estrogen-related receptor alpha is essential for adaptive thermogenesis. Proc. Natl Acad. Sci. USA 104, 1418–1423 (2007).

    CAS  PubMed  Google Scholar 

  80. Giguère, V. Transcriptional control of energy homeostasis by the estrogen-related receptors. Endocr. Rev. 29, 677–696 (2008).

    PubMed  Google Scholar 

  81. Luo, J. et al. Reduced fat mass in mice lacking orphan nuclear receptor estrogen-related receptor alpha. Mol. Cell. Biol. 23, 7947–7956 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  82. Chang, J. S., Ghosh, S., Newman, S. & Salbaum, J. M. A map of the PGC-1α- and NT-PGC-1α-regulated transcriptional network in brown adipose tissue. Sci. Rep. 8, 7876 (2018).

    PubMed  PubMed Central  Google Scholar 

  83. Schreiber, S. N., Knutti, D., Brogli, K., Uhlmann, T. & Kralli, A. The transcriptional coactivator PGC-1 regulates the expression and activity of the orphan nuclear receptor estrogen-related receptor alpha (ERRalpha). J. Biol. Chem. 278, 9013–9018 (2003).

    CAS  PubMed  Google Scholar 

  84. Brown, E. L. et al. Estrogen-related receptors mediate the adaptive response of brown adipose tissue to adrenergic stimulation. iScience 2, 221–237 (2018).

    PubMed  PubMed Central  Google Scholar 

  85. Ahmadian, M. et al. ERRγ preserves brown fat innate thermogenic activity. Cell Rep. 22, 2849–2859 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Seale, P. et al. Transcriptional control of brown fat determination by PRDM16. Cell Metab. 6, 38–54 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  87. Hasegawa, Y. et al. Repression of adipose tissue fibrosis through a PRDM16-GTF2IRD1 complex improves systemic glucose homeostasis. Cell Metab. 27, 180–194 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Berry, D. C. et al. Cellular aging contributes to failure of cold-induced beige adipocyte formation in old mice and humans. Cell Metab. 25, 481 (2017).

    CAS  PubMed  Google Scholar 

  89. Ferrari, A. et al. HDAC3 is a molecular brake of the metabolic switch supporting white adipose tissue browning. Nat. Commun. 8, 93 (2017).

    PubMed  PubMed Central  Google Scholar 

  90. Somech, R. et al. The nuclear-envelope protein and transcriptional repressor LAP2beta interacts with HDAC3 at the nuclear periphery, and induces histone H4 deacetylation. J. Cell Sci. 118, 4017–4025 (2005).

    CAS  PubMed  Google Scholar 

  91. Vincent, S. D. & Buckingham, M. E. How to make a heart. The origin and regulation of cardiac progenitor cells. Curr. Top. Dev. Biol. 90, 1–41 (2010).

    PubMed  Google Scholar 

  92. Lewandowski, S. L. et al. Histone deacetylase 3 modulates Tbx5 activity to regulate early cardiogenesis. Hum. Mol. Genet. 23, 3801–3809 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  93. Lewandowski, S. L., Janardhan, H. P. & Trivedi, C. M. Histone deacetylase 3 coordinates deacetylase-independent epigenetic silencing of transforming growth factor-β1 (TGF-β1) to orchestrate second heart field development. J. Biol. Chem. 290, 27067–27089 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Poleshko, A. et al. Genome-nuclear lamina interactions regulate cardiac stem cell lineage restriction. Cell 171, 573–587 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  95. Demmerle, J., Koch, A. J. & Holaska, J. M. The nuclear envelope protein emerin binds directly to histone deacetylase 3 (HDAC3) and activates HDAC3 activity. J. Biol. Chem. 287, 22080–22088 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  96. Zullo, J. M. et al. DNA sequence-dependent compartmentalization and silencing of chromatin at the nuclear lamina. Cell 149, 1474–1487 (2012).

    CAS  PubMed  Google Scholar 

  97. Jain, R. & Epstein, J. A. Competent for commitment: you’ve got to have heart! Genes Dev. 32, 4–13 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Van Der Linde, D. et al. Birth prevalence of congenital heart disease worldwide: a systematic review and meta-analysis. J. Am. Coll. Cardiol. 58, 2241–2247 (2011).

    PubMed  Google Scholar 

  99. Loeys, B. L. et al. A syndrome of altered cardiovascular, craniofacial, neurocognitive and skeletal development caused by mutations in TGFBR1 or TGFBR2. Nat. Genet. 37, 275–281 (2005).

    CAS  PubMed  Google Scholar 

  100. Neptune, E. R. et al. Dysregulation of TGF-beta activation contributes to pathogenesis in Marfan syndrome. Nat. Genet. 33, 407–411 (2003).

    CAS  PubMed  Google Scholar 

  101. Singh, N. et al. Histone deacetylase 3 regulates smooth muscle differentiation in neural crest cells and development of the cardiac outflow tract. Circ. Res. 109, 1240–1249 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  102. Sun, Z. et al. Diet-induced lethality due to deletion of the Hdac3 gene in heart and skeletal muscle. J. Biol. Chem. 286, 33301–33309 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  103. Vega, R. B. & Kelly, D. P. Cardiac nuclear receptors: architects of mitochondrial structure and function. J. Clin. Invest. 127, 1155–1164 (2017).

    PubMed  PubMed Central  Google Scholar 

  104. Finck, B. N. & Kelly, D. P. PGC-1 coactivators: Inducible regulators of energy metabolism in health and disease. J. Clin. Invest. 116, 615–622 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  105. 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 

  106. Zhang, L. et al. Hdac3 interaction with p300 histone acetyltransferase regulates the oligodendrocyte and astrocyte lineage fate switch. Dev. Cell 36, 316–330 (2016).

    PubMed  PubMed Central  Google Scholar 

  107. He, X. et al. A histone deacetylase 3-dependent pathway delimits peripheral myelin growth and functional regeneration. Nat. Med. 24, 338–351 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  108. 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 

  109. Rogge, G. A., Singh, H., Dang, R. & Wood, M. A. HDAC3 is a negative regulator of cocaine-context-associated memory formation. J. Neurosci. 33, 6623–6632 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  110. 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 

  111. 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 

  112. Nott, A. et al. Histone deacetylase 3 associates with MeCP2 to regulate FOXO and social behavior. Nat. Neurosci. 19, 1497–1505 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Wang, Y. et al. HDAC3-dependent epigenetic pathway controls lung alveolar epithelial cell remodeling and spreading via miR-17-92 and TGF-β signaling regulation. Dev. Cell 36, 303–315 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  114. Wang, X. et al. Expression of histone deacetylase 3 instructs alveolar type I cell differentiation by regulating a Wnt signaling niche in the lung. Dev. Biol. 414, 161–169 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  115. Alenghat, T. et al. Histone deacetylase 3 coordinates commensal-bacteria-dependent intestinal homeostasis. Nature 504, 153–157 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Navabi, N. et al. Epithelial histone deacetylase 3 instructs intestinal immunity by coordinating local lymphocyte activation. Cell Rep. 19, 1165–1175 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  117. Whitt, J. et al. Disruption of epithelial HDAC3 in intestine prevents diet-induced obesity in mice. Gastroenterology 155, 501–513 (2018).

    CAS  PubMed  Google Scholar 

  118. Remsberg, J. R. et al. Deletion of histone deacetylase 3 in adult beta cells improves glucose tolerance via increased insulin secretion. Mol. Metab. 6, 30–37 (2017).

    CAS  PubMed  Google Scholar 

  119. Lundh, M., Galbo, T., Poulsen, S. S. & Mandrup-Poulsen, T. Histone deacetylase 3 inhibition improves glycaemia and insulin secretion in obese diabetic rats. Diabetes Obes. Metab. 17, 703–707 (2015).

    CAS  PubMed  Google Scholar 

  120. Chou, D. H. C. et al. Inhibition of histone deacetylase 3 protects beta cells from cytokine-induced apoptosis. Chem. Biol. 19, 669–673 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  121. Lundh, M. et al. Histone deacetylases 1 and 3 but not 2 mediate cytokine-induced beta cell apoptosis in INS-1 cells and dispersed primary islets from rats and are differentially regulated in the islets of type 1 diabetic children. Diabetologia 55, 2421–2431 (2012).

    CAS  PubMed  Google Scholar 

  122. Chen, W.-B. et al. Conditional ablation of HDAC3 in islet beta cells results in glucose intolerance and enhanced susceptibility to STZ-induced diabetes. Oncotarget 7, 57485–57497 (2016).

    PubMed  PubMed Central  Google Scholar 

  123. Magnuson, M. A. & Osipovich, A. B. Pancreas-specific Cre driver lines and considerations for their prudent use. Cell Metab. 18, 9–20 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  124. Song, J., Xu, Y., Hu, X., Choi, B. & Tong, Q. Brain expression of Cre recombinase driven by pancreas-specific promoters. Genesis 48, 628–634 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  125. Hong, S. et al. Dissociation of muscle insulin sensitivity from exercise endurance in mice by HDAC3 depletion. Nat. Med. 23, 223–234 (2017).

    CAS  PubMed  Google Scholar 

  126. Mullican, S. E. et al. Histone deacetylase 3 is an epigenomic brake in macrophage alternative activation. Genes Dev. 25, 2480–2488 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  127. Chen, X. et al. Requirement for the histone deacetylase Hdac3 for the inflammatory gene expression program in macrophages. Proc. Natl Acad. Sci. USA 109, E2865–E2874 (2012).

    CAS  PubMed  Google Scholar 

  128. Hoeksema, M. a et al. Targeting macrophage Histone deacetylase 3 stabilizes atherosclerotic lesions. EMBO Mol. Med. 6, e201404170 (2014).

    Google Scholar 

  129. 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 

  130. Summers, A. R. et al. HDAC3 is essential for DNA replication in hematopoietic progenitor cells. J. Clin. Invest. 123, 3112–3123 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  131. Stengel, K. R. et al. Deacetylase activity of histone deacetylase 3 is required for productive VDJ recombination and B cell development. Proc. Natl Acad. Sci. USA 114, 8608–8613 (2017).

    CAS  PubMed  Google Scholar 

  132. Stengel, K. R. et al. Histone deacetylase 3 is required for efficient T cell development. Mol. Cell. Biol. 35, 3854–3865 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Philips, R. L. et al. HDAC3 is required for the downregulation of RORγt during thymocyte positive selection. J. Immunol. 197, 541–554 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  134. Hsu, F.-C. et al. Histone deacetylase 3 is required for T cell maturation. J. Immunol. 195, 1578–1590 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  135. Wang, L. et al. FOXP3 + regulatory T cell development and function require histone / protein deacetylase 3. J. Clin. Invest. 125, 1–13 (2015).

    Google Scholar 

  136. Chen Lf, Fischle, W., Verdin, E. & Greene, W. C. Duration of nuclear NF-kappaB action regulated by reversible acetylation. Science 293, 1653–1657 (2001).

    CAS  Google Scholar 

  137. Bradley, E. W., Carpio, L. R., van Wijnen, A. J., McGee-Lawrence, M. E. & Westendorf, J. J. Histone deacetylases in bone development and skeletal disorders. Physiol. Rev. 95, 1359–1381 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  138. Singh, N. et al. Murine craniofacial development requires Hdac3-mediated repression of Msx gene expression. Dev. Biol. 377, 333–344 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  139. Razidlo, D. F. et al. Histone deacetylase 3 depletion in osteo/chondroprogenitor cells decreases bone density and increases marrow fat. PLOS ONE 5, e11492 (2010).

    PubMed  PubMed Central  Google Scholar 

  140. Feigenson, M. et al. Histone deacetylase 3 deletion in mesenchymal progenitor cells hinders long bone development. J. Bone Miner. Res. 32, 2453–2465 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  141. Carpio, L. R. et al. Histone deacetylase 3 supports endochondral bone formation by controlling cytokine signaling and matrix remodeling. Sci. Signal. 9, ra79 (2016).

    PubMed  PubMed Central  Google Scholar 

  142. McGee-Lawrence, M. E. et al. Histone deacetylase 3 is required for maintenance of bone mass during aging. Bone 52, 296–307 (2013).

    CAS  PubMed  Google Scholar 

  143. Schroeder, T. M., Kahler, R. A., Li, X. & Westendorf, J. J. Histone deacetylase 3 interacts with Runx2 to repress the osteocalcin promoter and regulate osteoblast differentiation. J. Biol. Chem. 279, 41998–42007 (2004).

    CAS  PubMed  Google Scholar 

  144. McGee-Lawrence, M. E. et al. Runx2 protein represses axin2 expression in osteoblasts and is required for craniosynostosis in axin2-deficient mice. J. Biol. Chem. 288, 5291–5302 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  145. Hesse, E. et al. Zfp521 controls bone mass by HDAC3-dependent attenuation of Runx2 activity. J. Cell Biol. 191, 1271–1283 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  146. Giavini, E. & Menegola, E. Teratogenic activity of HDAC inhibitors. Curr. Pharm. Des. 20, 5438–5442 (2014).

    CAS  PubMed  Google Scholar 

  147. Boluk, A. et al. The effect of valproate on bone mineral density in adult epileptic patients. Pharmacol. Res. 50, 93–97 (2004).

    CAS  PubMed  Google Scholar 

  148. Guo, C. Y., Ronen, G. M. & Atkinson, S. A. Long-term valproate and lamotrigine treatment may be a marker for reduced growth and bone mass in children with epilepsy. Epilepsia 42, 1141–1147 (2001).

    CAS  PubMed  Google Scholar 

  149. Oner, N. et al. Bone mineral metabolism changes in epileptic children receiving valproic acid. J. Paediatr. Child Health 40, 470–473 (2004).

    CAS  PubMed  Google Scholar 

  150. Sato, Y. et al. Decreased bone mass and increased bone turnover with valproate therapy in adults with epilepsy. Neurology 57, 445–449 (2001).

    CAS  PubMed  Google Scholar 

  151. Lajeunie, E. et al. Craniosynostosis and fetal exposure to sodium valproate. J. Neurosurg. 95, 778–782 (2001).

    CAS  PubMed  Google Scholar 

  152. Sharony, R. et al. Preaxial ray reduction defects as part of valproic acid embryofetopathy. Prenat. Diagn. 13, 909–918 (1993).

    CAS  PubMed  Google Scholar 

  153. Paradis, F.-H. & Hales, B. F. The effects of class-specific histone deacetylase inhibitors on the development of limbs during organogenesis. Toxicol. Sci. 148, 220–228 (2015).

    CAS  PubMed  Google Scholar 

  154. Heinz, S. et al. Effect of natural genetic variation on enhancer selection and function. Nature 503, 487–492 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  155. Janardhan, H. P. et al. Hdac3 regulates lymphovenous and lymphatic valve formation. J. Clin. Invest. 127, 4193–4206 (2017).

    PubMed  PubMed Central  Google Scholar 

  156. Smith, E. & Shilatifard, A. Enhancer biology and enhanceropathies. Nat. Struct. Mol. Biol. 21, 210–219 (2014).

    CAS  PubMed  Google Scholar 

  157. Lerin, C. et al. GCN5 acetyltransferase complex controls glucose metabolism through transcriptional repression of PGC-1alpha. Cell Metab. 3, 429–438 (2006).

    CAS  PubMed  Google Scholar 

  158. Rodgers, J. T. et al. Nutrient control of glucose homeostasis through a complex of PGC-1alpha and SIRT1. Nature 434, 113–118 (2005).

    CAS  PubMed  Google Scholar 

  159. Wang, W. et al. Ebf2 is a selective marker of brown and beige adipogenic precursor cells. Proc. Natl Acad. Sci. USA 111, 14466–14471 (2014).

    CAS  PubMed  Google Scholar 

  160. Shapira, S. N. et al. EBF2 transcriptionally regulates brown adipogenesis via the histone reader DPF3 and the BAF chromatin remodeling complex. Genes Dev. 31, 660–673 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  161. Thompson, J. A., Tan, J. & Greene, C. S. Cross-platform normalization of microarray and RNA-seq data for machine learning applications. PeerJ 4, e1621 (2016).

    PubMed  PubMed Central  Google Scholar 

  162. Ritchie, M. E. et al. Limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 43, e47 (2015).

    PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank members of the Lazar laboratory for discussion and comments on the manuscript. They also thank H.-W. Lim for processing ChIP-seq and GRO-seq data from GEO data sets (GSE83928 and GSE110056), H. B. Nguyen for bioinformatics analysis of gene expression (GSE98650, GSE90531, GSE83927, GSE72917, GSE50188, GSE85929, GSE33609, GSE79696 and GSE68991) and H. J. Richter for assistance with modelling the crystal structure of HDAC3. US National Institutes of Health (NIH) grant R01 DK45586 (M.A.L.) and NIH F30 DK104513 (M.J.E.) supported this work.

Reviewer information

Nature Reviews Molecular Cell Biology thanks C. Trivedi and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Author information

Authors and Affiliations

Authors

Contributions

M.J.E. and M.A.L. researched data for the article, made substantial contributions to the writing and content and reviewed and edited the manuscript before submission.

Corresponding author

Correspondence to Mitchell A. Lazar.

Ethics declarations

Competing interests

M.A.L. is a consultant to KDAC, a company developing histone deacetylase (HDAC) inhibitors, and Novartis, and serves on scientific advisory boards for Pfizer and Eli Lilly and Co.

Additional information

Publisher’s note

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

Related links

Gene expression omnibus: https://www.ncbi.nlm.nih.gov/gds

Glossary

Lineage-determining transcription factors

Transcription factors that determine the differentiation fate of cells.

Signal-dependent transcription factors

Transcription factors that associate with or bind to specific DNA sequences near lineage-determining transcription factors in response to environmental or physiological cues.

Nuclear receptors

A superfamily of transcription factors that bind to highly specific DNA motifs in direct response to ligand binding to activate or repress gene transcription.

Stoichiometric component

A protein whose interaction with another component of a protein complex is based on their equal molarity.

WD40 repeat-containing proteins

Proteins containing a motif of 40 amino acids that assumes a β-propeller structure and functions in establishing protein–protein interactions, signal transduction and transcription regulation.

Cre recombinase

(Cre). A bacteriophage P1 type I topoisomerase that catalyses DNA recombination between site-specific loxP sites.

Adeno-associated virus

(AAV). A small, non-enveloped, replication-incompetent virus with a small single-stranded DNA genome, which is maintained extrachromosomally and used for gene delivery.

Hepatosteatosis

Fatty liver, often owing to the pathological accumulation of triglycerides.

Nonshivering thermogenesis

A facultative and adaptive process that protects core body temperature.

Oxidative phosphorylation

The transfer of electrons in mitochondria between energy carriers down the electron transport chain to generate a proton gradient and produce ATP.

Global run-on sequencing

(GRO–seq). A nuclear run-on assay coupled to next-generation sequencing to map all transcription by engaged RNA polymerases throughout the genome.

Futile metabolic cycles

Biochemical pathways that operate simultaneously in opposing directions, often leading to the dissipation of energy.

First heart field

(FHF). A developmental structure that gives rise to the cardiac crescent, early cardiac tube and left ventricle.

Second heart field

(SHF). A developmental structure that gives rise to the right ventricle, atrial myocardium and cardiac outflow tract.

Nuclear lamina

A cellular structure adjacent to the inner nuclear membrane that is composed of lamin polymers and other proteins and forms a skeletal nuclear structure that interacts with nuclear scaffold proteins and chromatin.

Facultative heterochromatin

Tightly packed, inaccessible chromatin containing inactive genes, which may become accessible and expressed in certain cell types or following certain cellular cues.

Marfan syndrome

A connective tissue disorder caused by mutations in the fibrillin 1 gene, often presenting as tall, slender individuals with arachnodactyly, cardiac valve disease and predisposition to aortic aneurysm and dissection.

Loeys–Dietz syndrome

A connective tissue disorder with five subtypes, each caused by unique mutations in five genes of the TGFβ signalling pathway, which cause congenital cardiac defects and predisposition to aortic aneurysm and dissection.

CA1 hippocampal region

An area of the brain that is important for memory.

Nucleus accumbens

A region of the brain that is involved in cognitive processing of reward, with neurons containing mostly dopamine receptors; the nucleus accumbens is thought to be the major nucleus involved in addiction.

Paneth cells

Epithelial cells of the small intestine that secrete antimicrobial peptides and proteins, mediate host–microorganism interactions and help defend against enteric pathogens in the gut lumen.

Regulatory T cells

(Treg cells). CD4+, CD25+ and FOXP3+ T cells that suppress the activation of the immune system to maintain tolerance to self-antigens and prevent autoimmune disease.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Emmett, M.J., Lazar, M.A. Integrative regulation of physiology by histone deacetylase 3. Nat Rev Mol Cell Biol 20, 102–115 (2019). https://doi.org/10.1038/s41580-018-0076-0

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41580-018-0076-0

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