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:

The genetics of human performance

Abstract

Human physiology is likely to have been selected for endurance physical activity. However, modern humans have become largely sedentary, with physical activity becoming a leisure-time pursuit for most. Whereas inactivity is a strong risk factor for disease, regular physical activity reduces the risk of chronic disease and mortality. Although substantial epidemiological evidence supports the beneficial effects of exercise, comparatively little is known about the molecular mechanisms through which these effects operate. Genetic and genomic analyses have identified genetic variation associated with human performance and, together with recent proteomic, metabolomic and multi-omic analyses, are beginning to elucidate the molecular genetic mechanisms underlying the beneficial effects of physical activity on human health.

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: VO2 max is affected by the function of numerous organs.
Fig. 2: The hypoxia-inducible factor signalling pathway affects the haematopoietic and vascular systems.
Fig. 3: Metabolic changes with acute exercise converge on PGC1α in skeletal muscle.
Fig. 4: Summary of molecular changes with acute exercise and long-term endurance training.

Similar content being viewed by others

References

  1. Carrier, D. R. et al. The energetic paradox of human running and hominid evolution [and comments and reply]. Curr. Anthropol. 25, 483–495 (1984).

    Google Scholar 

  2. Bramble, D. M. & Lieberman, D. E. Endurance running and the evolution of Homo. Nature 432, 345–352 (2004). This article popularized the ‘evolution for endurance’ hypothesis first proposed by Carrier et al. (1984).

    PubMed  CAS  Google Scholar 

  3. Crittenden A. N. & Marlowe F. W. in Attachment Reconsidered: Culture, Mind, and Society (eds Quinn N. & Mageo J. M.) (Palgrave Macmillan, 2013) Book Series of the Society for Psychological Anthropology.

  4. Schnohr, P., O’Keefe, J. H., Marott, J. L., Lange, P. & Jensen, G. B. Dose of jogging and long-term mortality: the Copenhagen City Heart Study. J. Am. Coll. Cardiol. 65, 411–419 (2015).

    PubMed  Google Scholar 

  5. Booth, F. W., Roberts, C. K., Thyfault, J. P., Ruegsegger, G. N. & Toedebusch, R. G. Role of inactivity in chronic diseases: evolutionary insight and pathophysiological mechanisms. Physiol. Rev. 97, 1351–1402 (2017).

    PubMed  PubMed Central  Google Scholar 

  6. Lieberman, D. E. Is exercise really medicine? An evolutionary perspective. Curr. Sport. Med. Rep. 14, 313–319 (2015).

    Google Scholar 

  7. Bouchard, C. et al. Familial aggregation ofV˙o 2 max response to exercise training: results from the HERITAGE family study. J. Appl. Physiol. 87, 1003–1008 (1999).

    PubMed  CAS  Google Scholar 

  8. de la Chapelle, A., Träskelin, A. L. & Juvonen, E. Truncated erythropoietin receptor causes dominantly inherited benign human erythrocytosis. Proc. Natl Acad. Sci. USA 90, 4495–4499 (1993). This article describes a study of the family of Eero Mäntyranta that identified a familial mutation that caused benign erythrocytosis and hence, improved VO2 max.

    PubMed  PubMed Central  Google Scholar 

  9. Sarzynski, M. A. & Bouchard, C. World-class athletic performance and genetic endowment. Nat. Metab. 2, 796–798 (2020).

    PubMed  Google Scholar 

  10. Verweij, N., Vegte, Y. J. V. D. & Harst, P. V. D. Genetic study links components of the autonomous nervous system to heart-rate profile during exercise. Nat. Commun. 9, 898 (2018).

    PubMed  PubMed Central  Google Scholar 

  11. Ramírez, J. et al. Thirty loci identified for heart rate response to exercise and recovery implicate autonomic nervous system. Nat. Commun. 9, 1947 (2018).

    PubMed  PubMed Central  Google Scholar 

  12. Bye, A. et al. Identification of novel genetic variants associated with cardiorespiratory fitness. Prog. Cardiovasc. Dis. 63, 341–349 (2020). This article describes the largest GWAS to date of VO2 max; it specifically highlights the role of an ACTN3 nonsense variant in optimizing EPA.

    PubMed  Google Scholar 

  13. Harvey, N. R. et al. Genetic variants associated with exercise performance in both moderately trained and highly trained individuals. Mol. Genet. Genomics 295, 515–523 (2020).

    PubMed  CAS  Google Scholar 

  14. Jacques, M. et al. Epigenetic changes in healthy human skeletal muscle following exercise – a systematic review. Epigenetics 14, 633–648 (2019). This excellent systematic review summarizes the evidence for epigenetic changes that occur with acute exercise.

    PubMed  PubMed Central  Google Scholar 

  15. Liu, X., Platt, C. & Rosenzweig, A. The role of microRNAs in the cardiac response to exercise. Cold Spring Harb. Perspect. Med. 7, a029850 (2017).

    PubMed  PubMed Central  Google Scholar 

  16. Bouchard, C. et al. Aims, design, and measurement protocol. Med. Sci. Sports Exerc. 27, 721 (1995). This article describes the original HERITAGE study, which laid the foundation for exercise genomics.

    PubMed  CAS  Google Scholar 

  17. Bouchard, C. et al. Familial resemblance for VO2max in the sedentary state: the HERITAGE family study. Med. Sci. Sports Exerc. 30, 252–258 (1998).

    PubMed  CAS  Google Scholar 

  18. Eynon, N., Morán, M., Birk, R. & Lucia, A. The champions’ mitochondria: is it genetically determined? A review on mitochondrial DNA and elite athletic performance. Physiol. Genomics 43, 789–798 (2011).

    PubMed  CAS  Google Scholar 

  19. Stefano, E., Marsigliante, S., Vetrugno, C. & Muscella, A. Is mitochondrial DNA profiling predictive for athletic performance? Mitochondrion 47, 125–138 (2019).

    PubMed  CAS  Google Scholar 

  20. Wone, B. W. M., Yim, W. C., Schutz, H., Meek, T. H. & Garland, T. Mitochondrial haplotypes are not associated with mice selectively bred for high voluntary wheel running. Mitochondrion 46, 134–139 (2019).

    PubMed  CAS  Google Scholar 

  21. Harvey, N. R. et al. Investigating the influence of mtDNA and nuclear encoded mitochondrial variants on high intensity interval training outcomes. Sci. Rep. 10, 11089 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  22. Timmons, J. A. et al. Using molecular classification to predict gains in maximal aerobic capacity following endurance exercise training in humans. J. Appl. Physiol. 108, 1487–1496 (2010).

    PubMed  PubMed Central  CAS  Google Scholar 

  23. Bouchard, C. et al. Genomic predictors of the maximal O2 uptake response to standardized exercise training programs. J. Appl. Physiol. 110, 1160–1170 (2011).

    PubMed  CAS  Google Scholar 

  24. Ghosh, S. et al. Integrative pathway analysis of a genome-wide association study of V˙O2max response to exercise training. J. Appl. Physiol. 115, 1343–1359 (2013).

    PubMed  PubMed Central  CAS  Google Scholar 

  25. Ahmetov, I. et al. Genome-wide association study identifies three novel genetic markers associated with elite endurance performance. Biol. Sport. 32, 3–9 (2015).

    PubMed  Google Scholar 

  26. Rankinen, T. et al. No evidence of a common DNA variant profile specific to world class endurance athletes. PLoS ONE 11, e0147330 (2016). Although a null GWAS on the ‘case status’ of elite endurance athletes, this study remains the largest and most multi-ethnic genetic study of variation affecting endurance performance.

    PubMed  PubMed Central  Google Scholar 

  27. Al-Khelaifi, F. et al. Genome-wide association study reveals a novel association between MYBPC3 gene polymorphism, endurance athlete status, aerobic capacity and steroid metabolism. Front. Genet. 11, 595 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  28. Papadimitriou, I. D. et al. No association between ACTN3 R577X and ACE I/D polymorphisms and endurance running times in 698 Caucasian athletes. BMC Genomics 19, 13 (2018).

    PubMed  PubMed Central  Google Scholar 

  29. Malczewska-Lenczowska, J. et al. HIF-1α and NFIA-AS2 polymorphisms as potential determinants of total hemoglobin mass in endurance athletes. J. Strength Cond. Res. https://doi.org/10.1519/JSC.0000000000003686 (2020).

    Article  PubMed  Google Scholar 

  30. Tadros, R. et al. Shared genetic pathways contribute to risk of hypertrophic and dilated cardiomyopathies with opposite directions of effect. Nat. Genet. 53, 128–134 (2021). This article describes a large GWAS of both cardiac function and dilated versus hypertrophic cardiomyopathy risk, which highlights the possibility of antagonist pleiotropy for genes that aid cardiac function during reproductive ages and then cause elevated risk later in life.

    PubMed  PubMed Central  CAS  Google Scholar 

  31. Kim, D. S., Gloyn, A. L. & Knowles, J. W. Genetics of type 2 diabetes: opportunities for precision medicine: JACC focus seminar. J. Am. Coll. Cardiol. 78, 496–512 (2021).

    PubMed  Google Scholar 

  32. Nezhad, F. Y. et al. Genes whose gain or loss-of-function increases endurance performance in mice: a systematic literature review. Front. Physiol. 10, 262 (2019).

    Google Scholar 

  33. Seto, J. T. et al. ACTN3 genotype influences muscle performance through the regulation of calcineurin signaling. J. Clin. Invest. 123, 4255–4263 (2013).

    PubMed  PubMed Central  CAS  Google Scholar 

  34. North, K. N. et al. A common nonsense mutation results in α-actinin-3 deficiency in the general population. Nat. Genet. 21, 353–354 (1999).

    PubMed  CAS  Google Scholar 

  35. Yang, N. et al. ACTN3 genotype is associated with human elite athletic performance. Am. J. Hum. Genet. 73, 627–631 (2003).

    PubMed  PubMed Central  CAS  Google Scholar 

  36. Coso, J. D. et al. More than a ‘speed gene’: ACTN3 R577X genotype, trainability, muscle damage, and the risk for injuries. Eur. J. Appl. Physiol. 119, 49–60 (2019). This study describes a mouse knockout model of ACTN3 that demonstrates fibre switching and enhanced endurance performance.

    PubMed  Google Scholar 

  37. MacArthur, D. G. et al. Loss of ACTN3 gene function alters mouse muscle metabolism and shows evidence of positive selection in humans. Nat. Genet. 39, 1261–1265 (2007).

    PubMed  CAS  Google Scholar 

  38. Lindholm, M. E. & Rundqvist, H. Skeletal muscle hypoxia-inducible factor-1 and exercise. Exp. Physiol. 101, 28–32 (2016).

    PubMed  CAS  Google Scholar 

  39. Lindholm, M. E. et al. Negative regulation of HIF in skeletal muscle of elite endurance athletes: a tentative mechanism promoting oxidative metabolism. Am. J. Physiol. Regul. Integr. Comp. Physiol. 307, R248–R255 (2014).

    PubMed  CAS  Google Scholar 

  40. Lorenzo, F. R. et al. A genetic mechanism for Tibetan high-altitude adaptation. Nat. Genet. 46, 951–956 (2014).

    PubMed  PubMed Central  CAS  Google Scholar 

  41. Mason, S. D. et al. Loss of skeletal muscle HIF-1α results in altered exercise endurance. PLoS Biol. 2, e288 (2004).

    PubMed  PubMed Central  Google Scholar 

  42. Wu, Z. et al. Mechanisms controlling mitochondrial biogenesis and respiration through the thermogenic coactivator PGC-1. Cell 98, 115–124 (1999).

    PubMed  CAS  Google Scholar 

  43. Kamei, Y. et al. PPARγ coactivator 1β/ERR ligand 1 is an ERR protein ligand, whose expression induces a high-energy expenditure and antagonizes obesity. Proc. Natl Acad. Sci. USA 100, 12378–12383 (2003).

    PubMed  PubMed Central  CAS  Google Scholar 

  44. Arany, Z. et al. Transcriptional coactivator PGC-1α controls the energy state and contractile function of cardiac muscle. Cell Metab. 1, 259–271 (2005).

    PubMed  CAS  Google Scholar 

  45. Gao, Z. et al. Butyrate Improves insulin sensitivity and increases energy expenditure in mice. Diabetes 58, 1509–1517 (2009).

    PubMed  PubMed Central  CAS  Google Scholar 

  46. Lantier, L. et al. AMPK controls exercise endurance, mitochondrial oxidative capacity, and skeletal muscle integrity. FASEB J. 28, 3211–3224 (2014).

    PubMed  CAS  Google Scholar 

  47. Michael, L. F. et al. Restoration of insulin-sensitive glucose transporter (GLUT4) gene expression in muscle cells by the transcriptional coactivator PGC-1. Proc. Natl Acad. Sci. USA 98, 3820–3825 (2001).

    PubMed  PubMed Central  CAS  Google Scholar 

  48. Fan, W. & Evans, R. M. Exercise mimetics: impact on health and performance. Cell Metab. 25, 242–247 (2017).

    PubMed  CAS  Google Scholar 

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

    PubMed  CAS  Google Scholar 

  50. Arany, Z. et al. HIF-independent regulation of VEGF and angiogenesis by the transcriptional coactivator PGC-1α. Nature 451, 1008–1012 (2008).

    PubMed  CAS  Google Scholar 

  51. Handschin, C. et al. Skeletal muscle fiber-type switching, exercise intolerance, and myopathy in PGC-1α muscle-specific knock-out animals. J. Biol. Chem. 282, 30014–30021 (2007).

    PubMed  CAS  Google Scholar 

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

    PubMed  CAS  Google Scholar 

  53. Eynon, N., Ruiz, J. R., Meckel, Y., Morán, M. & Lucia, A. Mitochondrial biogenesis related endurance genotype score and sports performance in athletes. Mitochondrion 11, 64–69 (2011).

    PubMed  CAS  Google Scholar 

  54. Semenova, E. A., Fuku, N. & Ahmetov, I. I. in Sports, Exercise, and Nutritional Genomics 73–104 (Elsevier, 2019)

  55. Persson, H. C., Danielsson, A. & Sunnerhagen, K. S. A cross sectional study of upper extremity strength ten days after a stroke; relationship between patient-reported and objective measures. BMC Neurol. 15, 178 (2015).

    PubMed  PubMed Central  Google Scholar 

  56. Bohannon, R. W. Test-retest reliability of measurements of hand-grip strength obtained by dynamometry from older adults: a systematic review of research in the pubmed database. J. Frailty Aging 6, 83–87 (2017).

    PubMed  CAS  Google Scholar 

  57. Leong, D. P. et al. Prognostic value of grip strength: findings from the Prospective Urban Rural Epidemiology (PURE) study. Lancet 386, 266–273 (2015).

    PubMed  Google Scholar 

  58. Reed, T., Fabsitz, R. R., Selby, J. V. & Carmelli, D. Genetic influences and grip strength norms in the NHLBI twin study males aged 59–69. Ann. Hum. Biol. 18, 425–432 (2009).

    Google Scholar 

  59. Tikkanen, E. et al. Biological insights into muscular strength: genetic findings in the UK biobank. Sci. Rep. 8, 6451 (2018). This article reports the largest GWAS to date of hand-grip strength that provides evidence for brain–muscle association.

    PubMed  PubMed Central  Google Scholar 

  60. Chan, J. P. L. et al. Genetics of hand grip strength in mid to late life. Age 37, 3 (2015).

    PubMed Central  Google Scholar 

  61. Matteini, A. M. et al. GWAS analysis of handgrip and lower body strength in older adults in the CHARGE consortium. Aging Cell 15, 792–800 (2016).

    PubMed  PubMed Central  CAS  Google Scholar 

  62. Willems, S. M. et al. Large-scale GWAS identifies multiple loci for hand grip strength providing biological insights into muscular fitness. Nat. Commun. 8, 16015 (2017).

    PubMed  PubMed Central  CAS  Google Scholar 

  63. SALE, D. G. Neural adaptation to resistance training. Med. Sci. Sports Exerc. 20, S135 (1988).

    PubMed  CAS  Google Scholar 

  64. Higginbotham, M. B. et al. Regulation of stroke volume during submaximal and maximal upright exercise in normal man. Circ. Res. 58, 281–291 (2018).

    Google Scholar 

  65. Vasan, R. S. et al. Genome-wide association of echocardiographic dimensions, brachial artery endothelial function and treadmill exercise responses in the Framingham Heart Study. BMC Med. Genet. 8, S2 (2007).

    PubMed  PubMed Central  Google Scholar 

  66. Vasan, R. S. et al. Genetic variants associated with cardiac structure and function: a meta-analysis and replication of genome-wide association data. JAMA 302, 168–178 (2009).

    PubMed  PubMed Central  CAS  Google Scholar 

  67. Fox, E. R. et al. Genome-wide association study of cardiac structure and systolic function in African Americans. Circ. Cardiovasc. Genet. 6, 37–46 (2018).

    Google Scholar 

  68. Wild, P. S. et al. Large-scale genome-wide analysis identifies genetic variants associated with cardiac structure and function. J. Clin. Invest. 127, 1798–1812 (2017).

    PubMed  PubMed Central  Google Scholar 

  69. Pirruccello, J. P. et al. Analysis of cardiac magnetic resonance imaging in 36,000 individuals yields genetic insights into dilated cardiomyopathy. Nat. Commun. 11, 2254 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  70. Kanai, M. et al. Genetic analysis of quantitative traits in the Japanese population links cell types to complex human diseases. Nat. Genet. 50, 390–400 (2018).

    PubMed  CAS  Google Scholar 

  71. Aung, N. et al. Genome-wide analysis of left ventricular image-derived phenotypes identifies fourteen loci associated with cardiac morphogenesis and heart failure development. Circulation 140, 1318–1330 (2019).

    PubMed  PubMed Central  CAS  Google Scholar 

  72. Choquet, H. et al. Meta-analysis of 26 638 individuals identifies two genetic loci associated with left ventricular ejection fraction. Circ. Genom. Precis. Med. 13, e002804 (2020).

    PubMed  PubMed Central  Google Scholar 

  73. Rovere, M. T. L. et al. Baroreflex sensitivity and heart-rate variability in prediction of total cardiac mortality after myocardial infarction. Lancet 351, 478–484 (1998).

    PubMed  Google Scholar 

  74. Ingelsson, E. et al. Heritability, linkage, and genetic associations of exercise treadmill test responses. Circulation 115, 2917–2924 (2007).

    PubMed  Google Scholar 

  75. RICE, T. et al. Heritability of HR and BP response to exercise training in the HERITAGE family study. Med. Sci. Sports Exerc. 34, 972–979 (2002).

    PubMed  Google Scholar 

  76. An, P. et al. Familial aggregation of exercise heart rate and blood pressure in response to 20 weeks of endurance training: the HERITAGE family study. Int. J. Sports Med. 24, 57–62 (2003).

    PubMed  CAS  Google Scholar 

  77. Rankinen, T. et al. Heritability of submaximal exercise heart rate response to exercise training is accounted for by nine SNPs. J. Appl. Physiol. 112, 892–897 (2012).

    PubMed  CAS  Google Scholar 

  78. van de Vegte, Y. J., Tegegne, B. S., Verweij, N., Snieder, H. & van der Harst, P. Genetics and the heart rate response to exercise. Cell Mol. Life Sci. 76, 2391–2409 (2019).

    PubMed  PubMed Central  Google Scholar 

  79. Eppinga, R. N. et al. Identification of genomic loci associated with resting heart rate and shared genetic predictors with all-cause mortality. Nat. Genet. 48, 1557–1563 (2016).

    PubMed  CAS  Google Scholar 

  80. Brodde, O. E. & Michel, M. C. Adrenergic and muscarinic receptors in the human heart. Pharmacol. Rev. 51, 651–690 (1999).

    PubMed  CAS  Google Scholar 

  81. Hautala, A. J. et al. Acetylcholine receptor M2 gene variants, heart rate recovery, and risk of cardiac death after an acute myocardial infarction. Ann. Med. 41, 197–207 (2009).

    PubMed  CAS  Google Scholar 

  82. MacDougall, J. D., Tuxen, D., Sale, D. G., Moroz, J. R. & Sutton, J. R. Arterial blood pressure response to heavy resistance exercise. J. Appl. Physiol. 58, 785–790 (1985).

    PubMed  CAS  Google Scholar 

  83. WILMORE, J. H. et al. Heart rate and blood pressure changes with endurance training: the HERITAGE family study. Med. Sci. Sport. Exer 33, 107–116 (2001).

    CAS  Google Scholar 

  84. Surendran, P. et al. Discovery of rare variants associated with blood pressure regulation through meta-analysis of 1.3 million individuals. Nat. Genet. 52, 1314–1332 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  85. Heinicke, K. et al. Blood volume and hemoglobin mass in elite athletes of different disciplines. Int. J. Sports Med. 22, 504–512 (2001).

    PubMed  CAS  Google Scholar 

  86. Lundby, C., Gassmann, M. & Pilegaard, H. Regular endurance training reduces the exercise induced HIF-1α and HIF-2α mRNA expression in human skeletal muscle in normoxic conditions. Eur. J. Appl. Physiol. 96, 363–369 (2006).

    PubMed  CAS  Google Scholar 

  87. Semenova, E. A. et al. The association of HFE gene H63D polymorphism with endurance athlete status and aerobic capacity: novel findings and a meta-analysis. Eur. J. Appl. Physiol. 120, 665–673 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  88. Ganesh, S. K. et al. Multiple loci influence erythrocyte phenotypes in the CHARGE consortium. Nat. Genet. 41, 1191–1198 (2009).

    PubMed  PubMed Central  CAS  Google Scholar 

  89. Vuckovic, D. et al. The polygenic and monogenic basis of blood traits and diseases. Cell 182, 1214–1231.e11 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  90. Chen, M.-H. et al. Trans-ethnic and ancestry-specific blood-cell genetics in 746,667 individuals from 5 global populations. Cell 182, 1198–1213.e14 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  91. Sarzynski, M. A., Ghosh, S. & Bouchard, C. Genomic and transcriptomic predictors of response levels to endurance exercise training. J. Physiol. 595, 2931–2939 (2017).

    PubMed  CAS  Google Scholar 

  92. Benton, M. L. et al. The influence of evolutionary history on human health and disease. Nat. Rev. Genet. 22, 269–283 (2021). This article is an excellent recent review of the role of evolution on gene selection, and how prior evolutionary pressures may affect disease risk in the modern era without selection pressures.

    PubMed  CAS  PubMed Central  Google Scholar 

  93. Bird, A. P. & Wolffe, A. P. Methylation-induced repression — belts, braces, and chromatin. Cell 99, 451–454 (1999).

    PubMed  CAS  Google Scholar 

  94. McGee, S. L. & Hargreaves, M. Epigenetics and exercise. Trends Endocrinol. Metab. 30, 636–645 (2019).

    PubMed  CAS  Google Scholar 

  95. McGee, S. L., Fairlie, E., Garnham, A. P. & Hargreaves, M. Exercise-induced histone modifications in human skeletal muscle. J. Physiol. 587, 5951–5958 (2009).

    PubMed  PubMed Central  CAS  Google Scholar 

  96. Barrès, R. et al. Acute exercise remodels promoter methylation in human skeletal muscle. Cell Metab. 15, 405–411 (2012). This seminal study provides much of the evidence for changes in DNA methylation after exercise.

    PubMed  Google Scholar 

  97. Lindholm, M. E. et al. An integrative analysis reveals coordinated reprogramming of the epigenome and the transcriptome in human skeletal muscle after training. Epigenetics 9, 1557–1569 (2015).

    Google Scholar 

  98. Theilen, N. T., Kunkel, G. H. & Tyagi, S. C. The role of exercise and TFAM in preventing skeletal muscle atrophy. J. Cell Physiol. 232, 2348–2358 (2017).

    PubMed  PubMed Central  CAS  Google Scholar 

  99. Pettersen, I. K. N. et al. Upregulated PDK4 expression is a sensitive marker of increased fatty acid oxidation. Mitochondrion 49, 97–110 (2019).

    PubMed  CAS  Google Scholar 

  100. Medrano, J. L. & Naya, F. J. The transcription factor MEF2A fine-tunes gene expression in the atrial and ventricular chambers of the adult heart. J. Biol. Chem. 292, 20975–20988 (2017).

    PubMed  PubMed Central  CAS  Google Scholar 

  101. Nitert, M. D. et al. Impact of an exercise intervention on DNA methylation in skeletal muscle from first-degree relatives of patients with type 2 diabetes. Diabetes 61, 3322–3332 (2012).

    PubMed  PubMed Central  CAS  Google Scholar 

  102. Turner, D. C. et al. DNA methylation across the genome in aged human skeletal muscle tissue and muscle-derived cells: the role of HOX genes and physical activity. Sci. Rep. 10, 15360 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  103. Pillon, N. J. et al. Transcriptomic profiling of skeletal muscle adaptations to exercise and inactivity. Nat. Commun. 11, 470 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  104. Amar, D. et al. Time trajectories in the transcriptomic response to exercise — a meta-analysis. Nat. Commun. 12, 3471 (2021). This article accompanies the online ExtraMeta (https://extrameta.org) resource, and highlights the changes to gene expression that occur with endurance exercise.

  105. Chapman, M. A. et al. Skeletal muscle transcriptomic comparison between long-term trained and untrained men and women. Cell Rep. 31, 107808 (2020). This study compares the transcriptome and molecular and physiological phenotypes of fit and trained endurance athletes with resistance athletes or untrained individuals.

    PubMed  CAS  Google Scholar 

  106. Contrepois, K. et al. Molecular choreography of acute exercise. Cell 181, 1112–1130.e16 (2020). This article compares the transcriptome and proteome of fit and trained endurance athletes with resistance athletes and untrained individuals.

    PubMed  PubMed Central  CAS  Google Scholar 

  107. Sanford, J. A. et al. Molecular transducers of physical activity consortium (MoTrPAC): mapping the dynamic responses to exercise. Cell 181, 1464–1474 (2020).

    PubMed  CAS  PubMed Central  Google Scholar 

  108. Pitsiladis, Y. P. et al. Athlome project consortium: a concerted effort to discover genomic and other “omic” markers of athletic performance. Physiol. Genomics 48, 183–190 (2016).

    PubMed  CAS  Google Scholar 

  109. Scheiman, J. et al. Meta-omics analysis of elite athletes identifies a performance-enhancing microbe that functions via lactate metabolism. Nat. Med. 25, 1104–1109 (2019). This article describes a longitudinal study with accompanying mouse evidence for the role of the microbiome in exercise performance.

    PubMed  PubMed Central  CAS  Google Scholar 

  110. Rabinowitz, J. D. & Enerbäck, S. Lactate: the ugly duckling of energy metabolism. Nat. Metab. 2, 566–571 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  111. Frampton, J., Murphy, K. G., Frost, G. & Chambers, E. S. Short-chain fatty acids as potential regulators of skeletal muscle metabolism and function. Nat. Metab. 2, 840–848 (2020).

    PubMed  CAS  Google Scholar 

  112. Schaum, N. et al. Ageing hallmarks exhibit organ-specific temporal signatures. Nature 583, 596–602 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  113. Vujkovic, M. et al. Discovery of 318 new risk loci for type 2 diabetes and related vascular outcomes among 1.4 million participants in a multi-ancestry meta-analysis. Nat. Genet. 52, 680–691 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  114. Eichler, E. E. et al. Missing heritability and strategies for finding the underlying causes of complex disease. Nat. Rev. Genet. 11, 446–450 (2010).

    PubMed  PubMed Central  CAS  Google Scholar 

  115. Landen, S. et al. Genetic and epigenetic sex-specific adaptations to endurance exercise. Epigenetics 14, 523–535 (2019).

    PubMed  PubMed Central  Google Scholar 

  116. Dikilitas, O. et al. Predictive utility of polygenic risk scores for coronary heart disease in three major racial and ethnic groups. Am. J. Hum. Genet. 106, 707–716 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  117. Fishman, D. et al. The effect of novel polymorphisms in the interleukin-6 (IL-6) gene on IL-6 transcription and plasma IL-6 levels, and an association with systemic-onset juvenile chronic arthritis. J. Clin. Invest. 102, 1369–1376 (1998).

    PubMed  PubMed Central  CAS  Google Scholar 

  118. Illig, T. et al. Significant association of the interleukin-6 gene polymorphisms C-174G and A-598G with Type 2 Diabetes. J. Clin. Endocrinol. Metab. 89, 5053–5058 (2004).

    PubMed  CAS  Google Scholar 

  119. Clément, K. et al. Genetic variation in the β3-adrenergic receptor and an increased capacity to gain weight in patients with morbid obesity. N. Engl. J. Med. 333, 352–354 (1995).

    PubMed  Google Scholar 

  120. Paré, G. et al. Genetic analysis of 103 candidate genes for coronary artery disease and associated phenotypes in a founder population reveals a new association between endothelin-1 and high-density lipoprotein cholesterol. Am. J. Hum. Genet. 80, 673–682 (2007).

    PubMed  PubMed Central  Google Scholar 

  121. Wang, Q. et al. Positional cloning of a novel potassium channel gene: KVLQT1 mutations cause cardiac arrhythmias. Nat. Genet. 12, 17–23 (1996).

    PubMed  Google Scholar 

  122. Chen, Y.-H. et al. KCNQ1 gain-of-function mutation in familial atrial fibrillation. Science 299, 251–254 (2003).

    PubMed  CAS  Google Scholar 

  123. Watkins, H. et al. Mutations in the cardiac myosin binding protein-C gene on chromosome 11 cause familial hypertrophic cardiomyopathy. Nat. Genet. 11, 434–437 (1995).

    PubMed  CAS  Google Scholar 

  124. Tannahill, G. M. et al. Succinate is an inflammatory signal that induces IL-1β through HIF-1α. Nature 496, 238–242 (2013).

    PubMed  PubMed Central  CAS  Google Scholar 

  125. Tomita, S. et al. Defective brain development in mice lacking the Hif-1α gene in neural cells. Mol. Cell Biol. 23, 6739–6749 (2003).

    PubMed  PubMed Central  CAS  Google Scholar 

  126. Wang, Y.-X. et al. Regulation of muscle fiber type and running endurance by PPARδ. PLoS Biol. 2, e294 (2004).

    PubMed  PubMed Central  Google Scholar 

  127. Narkar, V. A. et al. AMPK and PPARδ agonists are exercise mimetics. Cell 134, 405–415 (2008).

    PubMed  PubMed Central  CAS  Google Scholar 

  128. Patten, I. S. et al. Cardiac angiogenic imbalance leads to peripartum cardiomyopathy. Nature 485, 333–338 (2012).

    PubMed  PubMed Central  CAS  Google Scholar 

  129. Kong, X. et al. IRF4 is a key thermogenic transcriptional partner of PGC-1α. Cell 158, 69–83 (2014).

    PubMed  PubMed Central  CAS  Google Scholar 

  130. Egan, M. F. et al. The BDNF val66met polymorphism affects activity-dependent secretion of BDNF and human memory and hippocampal function. Cell 112, 257–269 (2003).

    PubMed  CAS  Google Scholar 

  131. Hall, D., Dhilla, A., Charalambous, A., Gogos, J. A. & Karayiorgou, M. Sequence variants of the brain-derived neurotrophic factor (BDNF) gene are strongly associated with obsessive-compulsive disorder. Am. J. Hum. Genet. 73, 370–376 (2003).

    PubMed  PubMed Central  CAS  Google Scholar 

  132. Soliman, F. et al. A genetic variant BDNF polymorphism alters extinction learning in both mouse and human. Science 327, 863–866 (2010).

    PubMed  PubMed Central  CAS  Google Scholar 

  133. Zhao, C. et al. Charcot-Marie-Tooth disease type 2A caused by mutation in a microtubule motor KIF1Bβ. Cell 105, 587–597 (2001).

    PubMed  CAS  Google Scholar 

  134. Gerull, B. et al. Mutations of TTN, encoding the giant muscle filament titin, cause familial dilated cardiomyopathy. Nat. Genet. 30, 201–204 (2002).

    PubMed  CAS  Google Scholar 

  135. Norton, N. et al. Genome-wide studies of copy number variation and exome sequencing identify rare variants in BAG3 as a cause of dilated cardiomyopathy. Am. J. Hum. Genet. 88, 273–282 (2011).

    PubMed  PubMed Central  CAS  Google Scholar 

  136. Wang, J. C. et al. Evidence of common and specific genetic effects: association of the muscarinic acetylcholine receptor M2 (CHRM2) gene with alcohol dependence and major depressive syndrome. Hum. Mol. Genet. 13, 1903–1911 (2004).

    PubMed  CAS  Google Scholar 

  137. Luo, X. et al. CHRM2 gene predisposes to alcohol dependence, drug dependence and affective disorders: results from an extended case–control structured association study. Hum. Mol. Genet. 14, 2421–2434 (2005).

    PubMed  CAS  Google Scholar 

  138. Ma, P. et al. The ubiquitin ligase RNF220 enhances canonical Wnt signaling through USP7-mediated deubiquitination of β-catenin. Mol. Cell Biol. 34, 4355–4366 (2014).

    PubMed  PubMed Central  Google Scholar 

  139. Song, D. et al. Tibetan PHD2, an allele with loss-of-function properties. Proc. Natl Acad. Sci. USA 117, 12230–12238 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  140. Percy, M. J. et al. A family with erythrocytosis establishes a role for prolyl hydroxylase domain protein 2 in oxygen homeostasis. Proc. Natl Acad. Sci. USA 103, 654–659 (2006).

    PubMed  PubMed Central  CAS  Google Scholar 

  141. Wood, B. & Collard, M. The human genus. Science 284, 65–71 (1999).

    PubMed  CAS  Google Scholar 

  142. AIello, L. & Dean, C. in An Introduction to Human Evolutionary Anatomy Ch. 15 (Elsevier, 2002) https://doi.org/10.1016/b978-0-08-057100-3.50019-5.

  143. Kram, R. & Taylor, C. R. Energetics of running: a new perspective. Nature 346, 265–267 (1990).

    PubMed  CAS  Google Scholar 

  144. Lieberman, D. E. Human locomotion and heat loss: an evolutionary perspective. Compr. Physiol. 5, 99–117 (2014).

    Google Scholar 

  145. Kamberov, Y. G. et al. Comparative evidence for the independent evolution of hair and sweat gland traits in primates. J. Hum. Evol. 125, 99–105 (2018).

    PubMed  PubMed Central  Google Scholar 

  146. Baccaredda-Boy, A. A note on the cutaneous arterial vessels of some primates. J. Vasc. Res. 1, 209–212 (1964).

    CAS  Google Scholar 

  147. Robertshaw, D. Sweat and heat exchange in man and other mammals. J. Hum. Evol. 14, 63–73 (1985).

    Google Scholar 

  148. Shave, R. E. et al. Selection of endurance capabilities and the trade-off between pressure and volume in the evolution of the human heart. Proc. Natl Acad. Sci. USA 116, 19905–19910 (2019).

    PubMed  PubMed Central  CAS  Google Scholar 

  149. Meyer, H. V. et al. Genetic and functional insights into the fractal structure of the heart. Nature 584, 589–594 (2020).

    PubMed  PubMed Central  CAS  Google Scholar 

  150. Minetti, A. E. Efficiency of equine express postal systems. Nature 426, 785–786 (2003).

    PubMed  CAS  Google Scholar 

  151. Lieberman, D. E. et al. Foot strike patterns and collision forces in habitually barefoot versus shod runners. Nature 463, 531–535 (2010).

    PubMed  CAS  Google Scholar 

  152. Schmidt-Nielsen, K., Bretz, W. L. & Taylor, C. R. Panting in dogs: unidirectional air flow over evaporative surfaces. Science 169, 1102–1104 (1970).

    PubMed  CAS  Google Scholar 

  153. Taylor, C. & Rowntree, V. Temperature regulation and heat balance in running cheetahs: a strategy for sprinters? Am. J. Physiol. 224, 848–851 (1973).

    PubMed  CAS  Google Scholar 

  154. Liebenberg, L. P. Hunting by modern hunter-gatherers. Curr. Anthropol. 47, 1017–1026 (2006).

    Google Scholar 

Download references

Acknowledgements

The authors thank J. W. Knowles, M. E. Lindholm and C. M. Mattsson for their feedback on drafts of this article. D.S.K. thanks L. Han, N. Liu, M. Roy-O’Reilly and G. Salzman for excellent patient care, allowing for the time needed to write this article. D.S.K. also thanks J. K. Rathkey for stimulating conversations on biomechanical efficiency and exercise metabolism.

Author information

Authors and Affiliations

Authors

Contributions

E.A.A. and D.S.K. wrote the article. All authors researched data for the article, had substantial contribution to the discussion of the content, and reviewed and edited the manuscript before submission.

Corresponding author

Correspondence to Euan A. Ashley.

Ethics declarations

Competing interests

M.T.W. reports grants and personal fees from Verily, Myokardia and ArrayBio, and consultancy fees from BioTelemetry, outside the submitted work. E.A.A. reports advisory board fees from Apple and Foresite Labs. E.A.A. has ownership interest in Nuevocor, DeepCell and Personalis, outside the submitted work. E.A.A. declares ownership interest in SVEXA, and is a board member of AstraZeneca. D.S.K. declares no competing interests.

Additional information

Peer review information

Nature Reviews Genetics thanks I. Ahmetov, N. Fuku and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note

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

Related links

Athlome Project Consortium: http://www.athlomeconsortium.org/

ExtraMeta: http://www.extrameta.org

Exercise at the Limit — Inherited Traits of Endurance: https://med.stanford.edu/elite.html

MetaMEx: http://www.metamex.eu

Molecular Transducers of Physical Activity Consortium: https://motrpac-data.org

Glossary

Endurance physical activity

(EPA). Human performance that optimizes sustained activity over an extended period of time, such as running, cycling and cross-country skiing. Maximal oxygen uptake (VO2 max) is commonly used as a proxy for maximal capacity for EPA.

Resistance physical activity

(RPA). Human performance that requires muscles to work against an external force, such as weightlifting; individual muscle groups (such as biceps or triceps) can be measured accurately in isolation using dynamometry.

Healthspan

The length of time that a person is healthy and independent in their activities; by contrast, lifespan simply measures the length of time a person is alive but does not factor in quality of life.

Maximal oxygen uptake

(VO2 max). Represents the maximum capacity for endurance physical activity (EPA). It is the maximal rate of oxygen uptake (O2 in ml kg−1 min−1) measured during incremental increases in EPA activity intensity, commonly on a treadmill or stationary bike.

Haplogroups

Combinations of alleles at different regions of the genome (organellar or nuclear) that share polymorphisms inherited from a common ancestor.

Polygenic score

A measure constructed from additive genotypes of variants across multiple genes associated with a trait that can be used to predict from an individual’s genetic sequence their likelihood of having that trait.

Type II/fast-twitch muscle fibre

This muscle fibre type is optimized for resistance physical activity, and is characterized by high force generation but low endurance capability.

Type I/oxidative muscle fibre

This ‘slow-twitch oxidative’ muscle fibre type is optimized for endurance physical activity; it has greater mitochondrial density, which leads to lower force generation but increased resistance to fatigue.

Expression quantitative trait locus

(eQTL). A genetic variant that is associated with a gene expression phenotype.

Cardiac output

The amount of blood the heart pumps through the circulatory system (measured in litres per minute and estimated by multiplying the heart rate by stroke volume). Cardiac output is one of the primary contributors to VO2 max.

Antagonistic pleiotropy

Occurs when a locus or variant has a beneficial effect on one trait but confers deleterious effects on a separate trait.

Epigenetic modifications

Chemical alterations to DNA (most commonly by methylation) or DNA-associated histones (via acetylation, phosphorylation or methylation) that result in heritable changes to gene expression without changing the DNA sequence.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kim, D.S., Wheeler, M.T. & Ashley, E.A. The genetics of human performance. Nat Rev Genet 23, 40–54 (2022). https://doi.org/10.1038/s41576-021-00400-5

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41576-021-00400-5

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