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:

Dissociation of muscle insulin sensitivity from exercise endurance in mice by HDAC3 depletion

Abstract

Type 2 diabetes and insulin resistance are associated with reduced glucose utilization in the muscle and poor exercise performance. Here we find that depletion of the epigenome modifier histone deacetylase 3 (HDAC3) specifically in skeletal muscle causes severe systemic insulin resistance in mice but markedly enhances endurance and resistance to muscle fatigue, despite reducing muscle force. This seemingly paradoxical phenotype is due to lower glucose utilization and greater lipid oxidation in HDAC3-depleted muscles, a fuel switch caused by the activation of anaplerotic reactions driven by AMP deaminase 3 (Ampd3) and catabolism of branched-chain amino acids. These findings highlight the pivotal role of amino acid catabolism in muscle fatigue and type 2 diabetes pathogenesis. Further, as genome occupancy of HDAC3 in skeletal muscle is controlled by the circadian clock, these results delineate an epigenomic regulatory mechanism through which the circadian clock governs skeletal muscle bioenergetics. These findings suggest that physical exercise at certain times of the day or pharmacological targeting of HDAC3 could potentially be harnessed to alter systemic fuel metabolism and exercise performance.

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: Deletion of HDAC3 in skeletal muscle reduces glucose uptake and insulin sensitivity independently of the upstream insulin signaling.
Figure 2: HDAC3 deletion reduces glucose utilization during muscle contractions but enhances exercise endurance and oxidative metabolism.
Figure 3: HDAC3 regulates muscle fuel preference and controls amino acid metabolism.
Figure 4: Enhanced amino acid catabolism underlies the glucose-to-lipid fuel switch in HDAC3-depleted muscles.
Figure 5: Nonbiased identification of the circadian clock as an upstream regulator of muscle HDAC3.
Figure 6: HDAC3 couples circadian cues with the regulation of genes in anaplerotic reactions and amino acid catabolism.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

References

  1. DeFronzo, R.A. & Tripathy, D. Skeletal muscle insulin resistance is the primary defect in type 2 diabetes. Diabetes Care 32 (Suppl. 2) S157–S163 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Samuel, V.T. & Shulman, G.I. Mechanisms for insulin resistance: common threads and missing links. Cell 148, 852–871 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Dubé, J.J. et al. Exercise-induced alterations in intramyocellular lipids and insulin resistance: the athlete's paradox revisited. Am. J. Physiol. Endocrinol. Metab. 294, E882–E888 (2008).

    PubMed  PubMed Central  Google Scholar 

  4. Brüning, J.C. et al. A muscle-specific insulin receptor knockout exhibits features of the metabolic syndrome of NIDDM without altering glucose tolerance. Mol. Cell 2, 559–569 (1998).

    PubMed  Google Scholar 

  5. Kim, J.K. et al. Redistribution of substrates to adipose tissue promotes obesity in mice with selective insulin resistance in muscle. J. Clin. Invest. 105, 1791–1797 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Holloszy, J.O. “Deficiency” of mitochondria in muscle does not cause insulin resistance. Diabetes 62, 1036–1040 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Lowell, B.B. & Shulman, G.I. Mitochondrial dysfunction and type 2 diabetes. Science 307, 384–387 (2005).

    CAS  PubMed  Google Scholar 

  8. Szendroedi, J., Phielix, E. & Roden, M. The role of mitochondria in insulin resistance and type 2 diabetes mellitus. Nat. Rev. Endocrinol. 8, 92–103 (2011).

    PubMed  Google Scholar 

  9. Muoio, D.M. & Neufer, P.D. Lipid-induced mitochondrial stress and insulin action in muscle. Cell Metab. 15, 595–605 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Randle, P.J., Garland, P.B., Hales, C.N. & Newsholme, E.A. The glucose fatty-acid cycle. Its role in insulin sensitivity and the metabolic disturbances of diabetes mellitus. Lancet 1, 785–789 (1963).

    CAS  PubMed  Google Scholar 

  11. Hue, L. & Taegtmeyer, H. The Randle cycle revisited: a new head for an old hat. Am. J. Physiol. Endocrinol. Metab. 297, E578–E591 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Kelley, D.E. & Mandarino, L.J. Fuel selection in human skeletal muscle in insulin resistance: a reexamination. Diabetes 49, 677–683 (2000).

    CAS  PubMed  Google Scholar 

  13. Muoio, D.M. Metabolic inflexibility: when mitochondrial indecision leads to metabolic gridlock. Cell 159, 1253–1262 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Zhang, R., Lahens, N.F., Ballance, H.I., Hughes, M.E. & Hogenesch, J.B. A circadian gene expression atlas in mammals: implications for biology and medicine. Proc. Natl. Acad. Sci. USA 111, 16219–16224 (2014).

    CAS  PubMed  Google Scholar 

  15. Kornmann, B., Schaad, O., Bujard, H., Takahashi, J.S. & Schibler, U. System-driven and oscillator-dependent circadian transcription in mice with a conditionally active liver clock. PLoS Biol. 5, e34 (2007).

    PubMed  PubMed Central  Google Scholar 

  16. Hatori, M. et al. Time-restricted feeding without reducing caloric intake prevents metabolic diseases in mice fed a high-fat diet. Cell Metab. 15, 848–860 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Bass, J. & Takahashi, J.S. Circadian integration of metabolism and energetics. Science 330, 1349–1354 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Yin, L. & Lazar, M.A. The orphan nuclear receptor Rev-erbalpha recruits the N-CoR/histone deacetylase 3 corepressor to regulate the circadian Bmal1 gene. Mol. Endocrinol. 19, 1452–1459 (2005).

    CAS  PubMed  Google Scholar 

  19. 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 

  20. 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 

  21. Sun, Z., Feng, D., Everett, L.J., Bugge, A. & Lazar, M.A. Circadian epigenomic remodeling and hepatic lipogenesis: lessons from HDAC3. Cold Spring Harb. Symp. Quant. Biol. 76, 49–55 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Harfmann, B.D., Schroder, E.A. & Esser, K.A. Circadian rhythms, the molecular clock, and skeletal muscle. J. Biol. Rhythms 30, 84–94 (2015).

    CAS  PubMed  Google Scholar 

  23. Mayeuf-Louchart, A., Staels, B. & Duez, H. Skeletal muscle functions around the clock. Diabetes Obes. Metab. 17 (Suppl. 1), 39–46 (2015).

    CAS  PubMed  Google Scholar 

  24. Kang, S., Tsai, L.T.-Y. & Rosen, E.D. Nuclear Mechanisms of Insulin Resistance. Trends Cell Biol. 26, 341–351 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Bottomley, M.J. et al. Structural and functional analysis of the human HDAC4 catalytic domain reveals a regulatory structural zinc-binding domain. J. Biol. Chem. 283, 26694–26704 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Fischle, W. et al. Enzymatic activity associated with class II HDACs is dependent on a multiprotein complex containing HDAC3 and SMRT/N-CoR. Mol. Cell 9, 45–57 (2002).

    CAS  PubMed  Google Scholar 

  27. 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 

  28. Bothe, G.W., Haspel, J.A., Smith, C.L., Wiener, H.H. & Burden, S.J. Selective expression of Cre recombinase in skeletal muscle fibers. Genesis 26, 165–166 (2000).

    CAS  PubMed  Google Scholar 

  29. Loro, E. et al. IL-15Rα is a determinant of muscle fuel utilization, and its loss protects against obesity. Am. J. Physiol. Regul. Integr. Comp. Physiol. 309, R835–R844 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Zierath, J.R. & Hawley, J.A. Skeletal muscle fiber type: influence on contractile and metabolic properties. PLoS Biol. 2, e348 (2004).

    PubMed  PubMed Central  Google Scholar 

  31. Schiaffino, S. & Reggiani, C. Fiber types in mammalian skeletal muscles. Physiol. Rev. 91, 1447–1531 (2011).

    CAS  PubMed  Google Scholar 

  32. Nedachi, T., Fujita, H. & Kanzaki, M. Contractile C2C12 myotube model for studying exercise-inducible responses in skeletal muscle. Am. J. Physiol. Endocrinol. Metab. 295, E1191–E1204 (2008).

    CAS  PubMed  Google Scholar 

  33. Fang, B. et al. Circadian enhancers coordinate multiple phases of rhythmic gene transcription in vivo. Cell 159, 1140–1152 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Lin, J. et al. Transcriptional co-activator PGC-1 alpha drives the formation of slow-twitch muscle fibres. Nature 418, 797–801 (2002).

    CAS  PubMed  Google Scholar 

  35. Calvo, J.A. et al. Muscle-specific expression of PPARgamma coactivator-1alpha improves exercise performance and increases peak oxygen uptake. J. Appl. Physiol. 104, 1304–1312 (2008).

    CAS  PubMed  Google Scholar 

  36. Choi, C.S. et al. Paradoxical effects of increased expression of PGC-1alpha on muscle mitochondrial function and insulin-stimulated muscle glucose metabolism. Proc. Natl. Acad. Sci. USA 105, 19926–19931 (2008).

    CAS  PubMed  Google Scholar 

  37. Wong, K.E. et al. Muscle-Specific Overexpression of PGC-1α Does Not Augment Metabolic Improvements in Response to Exercise and Caloric Restriction. Diabetes 64, 1532–1543 (2015).

    CAS  PubMed  Google Scholar 

  38. Finck, B.N. et al. A potential link between muscle peroxisome proliferator- activated receptor-alpha signaling and obesity-related diabetes. Cell Metab. 1, 133–144 (2005).

    CAS  PubMed  Google Scholar 

  39. Pagel-Langenickel, I., Bao, J., Pang, L. & Sack, M.N. The role of mitochondria in the pathophysiology of skeletal muscle insulin resistance. Endocr. Rev. 31, 25–51 (2010).

    CAS  PubMed  Google Scholar 

  40. 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 

  41. Woldt, E. et al. Rev-erb-α modulates skeletal muscle oxidative capacity by regulating mitochondrial biogenesis and autophagy. Nat. Med. 19, 1039–1046 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Yamamoto, H. et al. NCoR1 is a conserved physiological modulator of muscle mass and oxidative function. Cell 147, 827–839 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Lau, P., Fitzsimmons, R.L., Pearen, M.A., Watt, M.J. & Muscat, G.E.O. Homozygous staggerer (sg/sg) mice display improved insulin sensitivity and enhanced glucose uptake in skeletal muscle. Diabetologia 54, 1169–1180 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Samant, S.A., Pillai, V.B., Sundaresan, N.R., Shroff, S.G. & Gupta, M.P. Histone Deacetylase 3 (HDAC3)-dependent Reversible Lysine Acetylation of Cardiac Myosin Heavy Chain Isoforms Modulates Their Enzymatic and Motor Activity. J. Biol. Chem. 290, 15559–15569 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Wagenmakers, A.J. Muscle amino acid metabolism at rest and during exercise: role in human physiology and metabolism. Exerc. Sport Sci. Rev. 26, 287–314 (1998).

    CAS  PubMed  Google Scholar 

  46. Gibala, M.J., Young, M.E. & Taegtmeyer, H. Anaplerosis of the citric acid cycle: role in energy metabolism of heart and skeletal muscle. Acta Physiol. Scand. 168, 657–665 (2000).

    CAS  PubMed  Google Scholar 

  47. Arinze, I.J. Facilitating understanding of the purine nucleotide cycle and the one-carbon pool: Part I: The purine nucleotide cycle. Biochem. Mol. Biol. Educ. 33, 165–168 (2005).

    CAS  PubMed  Google Scholar 

  48. Bowtell, J.L., Marwood, S., Bruce, M., Constantin-Teodosiu, D. & Greenhaff, P.L. Tricarboxylic acid cycle intermediate pool size: functional importance for oxidative metabolism in exercising human skeletal muscle. Sports Med. 37, 1071–1088 (2007).

    PubMed  Google Scholar 

  49. She, P. et al. Disruption of BCAA metabolism in mice impairs exercise metabolism and endurance. J. Appl. Physiol. 108, 941–949 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Ginevičienė, V. et al. AMPD1 rs17602729 is associated with physical performance of sprint and power in elite Lithuanian athletes. BMC Genet. 15, 58 (2014).

    PubMed  PubMed Central  Google Scholar 

  51. Van den Berghe, G., Bontemps, F., Vincent, M.F. & Van den Bergh, F. The purine nucleotide cycle and its molecular defects. Prog. Neurobiol. 39, 547–561 (1992).

    CAS  PubMed  Google Scholar 

  52. Operti, M.G., Vincent, M.F., Brucher, J.M. & van den Berghe, G. Enzymes of the purine nucleotide cycle in muscle of patients with exercise intolerance. Muscle Nerve 21, 401–403 (1998).

    CAS  PubMed  Google Scholar 

  53. Templeman, N.M., Schutz, H., Garland, T. Jr. & McClelland, G.B. Do mice bred selectively for high locomotor activity have a greater reliance on lipids to power submaximal aerobic exercise? Am. J. Physiol. Regul. Integr. Comp. Physiol. 303, R101–R111 (2012).

    CAS  PubMed  Google Scholar 

  54. Overmyer, K.A. et al. Maximal oxidative capacity during exercise is associated with skeletal muscle fuel selection and dynamic changes in mitochondrial protein acetylation. Cell Metab. 21, 468–478 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. She, P. et al. Disruption of BCATm in mice leads to increased energy expenditure associated with the activation of a futile protein turnover cycle. Cell Metab. 6, 181–194 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Cheng, J. et al. AMPD1: a novel therapeutic target for reversing insulin resistance. BMC Endocr. Disord. 14, 96 (2014).

    PubMed  PubMed Central  Google Scholar 

  57. Jang, C. et al. A branched-chain amino acid metabolite drives vascular fatty acid transport and causes insulin resistance. Nat. Med. 22, 421–426 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Hawley, J.A., Schabort, E.J., Noakes, T.D. & Dennis, S.C. Carbohydrate-loading and exercise performance. An update. Sports Med. 24, 73–81 (1997).

    CAS  PubMed  Google Scholar 

  59. Phinney, S.D. Ketogenic diets and physical performance. Nutr. Metab. (Lond.) 1, 2 (2004).

    Google Scholar 

  60. Paoli, A., Bianco, A. & Grimaldi, K.A. The Ketogenic Diet and Sport: A Possible Marriage? Exerc. Sport Sci. Rev. 43, 153–162 (2015).

    PubMed  Google Scholar 

  61. Cheng, R. et al. Efficient gene editing in adult mouse livers via adenoviral delivery of CRISPR/Cas9. FEBS Lett. 588, 3954–3958 (2014).

    CAS  PubMed  Google Scholar 

  62. Lu, W. et al. Metabolomic analysis via reversed-phase ion-pairing liquid chromatography coupled to a stand alone orbitrap mass spectrometer. Anal. Chem. 82, 3212–3221 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Melamud, E., Vastag, L. & Rabinowitz, J.D. Metabolomic analysis and visualization engine for LC-MS data. Anal. Chem. 82, 9818–9826 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Sun, Z. et al. Deacetylase-independent function of HDAC3 in transcription and metabolism requires nuclear receptor corepressor. Mol. Cell 52, 769–782 (2013).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank S. Burden (New York University) for MLC-Cre mice, J. Hogenesch (University of Pennsylvania) for processed data from circaDB, M. Birnbaum and J. Baur (University of Pennsylvania) for helpful discussion, S. Guo (Texas A&M University) for help with the EPS procedure, P. Zhang (Baylor College of Medicine) for adenoviral plasmids, S. Hui and J. Park (Princeton University) for analysis of metabolomics data, V. Narkar (University of Texas Health Science Center) for the immunostaining protocol, M. Goncalves and C. Lanzillotta (University of Pennsylvania) for technical assistance. We thank the Penn Diabetes Center (DK19525) Functional Genomics Core for nucleotide sequencing, the Mouse Metabolic Phenotyping Core for clamp experiments, the Penn Muscle Institute Muscle Physiology Assessment Core for muscle contraction study, and the Princeton/Penn Regional Metabolomics Core for flux and lipid analysis. We thank the Baylor Diabetes Center (DK079638) Metabolism Core for Seahorse analysis and Vanderbilt MMPC (DK59637) for lipidomics analysis. The Rockefeller Proteomics Resource Center is supported by the Leona M. and Harry B. Helmsley Charitable Trust. This work was supported by NIH grants CA211437 (W.L.), DK043806 (M.A.L.) and DK099443 (Z.S.).

Author information

Authors and Affiliations

Authors

Contributions

Z.S. and M.A.L. conceived the study and designed experiments. S.H., W.Z., W.L., E.L., G.D., J.J., S.Z., Y.Z., D.F., Q.C., B.D.D. and Z.S. conducted experiments. S.H., W.Z., B.F., W.L., M.D., G.D., S.Z., B.D.D., H.M. and Z.S. analyzed the data. E.L., T.S.K., J.D.R., M.A.L. and Z.S. interpreted the data. M.A.L. and Z.S. acquired funding. Z.S. wrote the manuscript with input from other authors.

Corresponding authors

Correspondence to Mitchell A Lazar or Zheng Sun.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Supplementary information

Supplementary Figures

Supplementary Figures 1–7 (PDF 2216 kb)

Supplementary Table 1

Raw data of metabolomics study. (XLSX 27 kb)

Supplementary Table 2

Differentially expressed genes from RNA-seq. (XLSX 44 kb)

Supplementary Table 3

Differentially expressed proteins from proteomics. (XLSX 475 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hong, S., Zhou, W., Fang, B. et al. Dissociation of muscle insulin sensitivity from exercise endurance in mice by HDAC3 depletion. Nat Med 23, 223–234 (2017). https://doi.org/10.1038/nm.4245

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nm.4245

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