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:

Pharmacological approaches to restore mitochondrial function

Key Points

  • Over the past decade, more evidence has accumulated regarding the involvement of mitochondrial dysfunction in the pathogenesis of type 2 diabetes, sarcopaenia, Parkinson's disease, cancer and, to a lesser extent, Alzheimer's disease. Modulating mitochondrial function has therefore emerged as an attractive strategy to treat both inherited mitochondrial diseases and age-related diseases associated with mitochondrial dysfunction, and spurred active drug discovery efforts in this area.

  • Several pathways are important to maintain and/or restore mitochondrial function and hence have potential therapeutic value. The best-characterized pathway is that involving the regulation of mitochondrial function by sensors of the nutrient or energy status, such as the pathway involving AMP-activated protein kinase (AMPK) and sirtuin 1 (SIRT1) and their downstream effectors. Other key pathways that regulate mitochondrial function, such as fusion and/or fission, mitophagy and the mitochondrial unfolded protein response, have been recently added to the equation, but there are very few compounds available to specifically modulate these pathways.

  • Most of the compounds that modulate mitochondrial function have been identified via target-based screens, including compounds that bind to and modulate G-protein-coupled receptors (such as G protein-coupled bile acid receptor 1), nuclear receptors and other transcription factors (such as peroxisome proliferator-activated receptors and oestrogen-related receptors), kinases (such as AMPK) and enzymes (such as SIRT1).

  • Recently, several phenotypic screening strategies have enabled the identification of novel chemical entities that modulate more global phenotypes, such as fusion and mitochondrial biogenesis. These kinds of strategies can be a powerful approach for identifying novel pathways or proteins involved in the regulatory network of mitochondrial function.

  • By combining the results of several phenotypic tests, it is possible to define a comprehensive set of mitochondrial signatures that enable the clustering of novel chemical entities according to their footprint on mitochondrial function. Such a signature will also help determine the particular mitochondrial signalling pathway that is affected, facilitating downstream target deconvolution.

Abstract

Mitochondrial dysfunction is not only a hallmark of rare inherited mitochondrial disorders but also implicated in age-related diseases, including those that affect the metabolic and nervous system, such as type 2 diabetes and Parkinson's disease. Numerous pathways maintain and/or restore proper mitochondrial function, including mitochondrial biogenesis, mitochondrial dynamics, mitophagy and the mitochondrial unfolded protein response. New and powerful phenotypic assays in cell-based models as well as multicellular organisms have been developed to explore these different aspects of mitochondrial function. Modulating mitochondrial function has therefore emerged as an attractive therapeutic strategy for several diseases, which has spurred active drug discovery efforts in this area.

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: Pharmacological approaches for targeting mitochondrial biogenesis.
Figure 2: Mitochondrial quality control processes.
Figure 3: Current models of mitochondrial quality control processes.
Figure 4: Models and approaches for mitochondrial screens.

Similar content being viewed by others

References

  1. Wallin, I. E. Symbionticism and the Origin of Species (Williams & Wilkins Company, 1927).

    Book  Google Scholar 

  2. Pagliarini, D. J. et al. A mitochondrial protein compendium elucidates complex I disease biology. Cell 134, 112–123 (2008). This study provides the list of mitochondrial proteins.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Ryan, M. T. & Hoogenraad, N. J. Mitochondrial-nuclear communications. Annu. Rev. Biochem. 76, 701–722 (2007).

    Article  CAS  PubMed  Google Scholar 

  4. Haynes, C. M. & Ron, D. The mitochondrial UPR — protecting organelle protein homeostasis. J. Cell Sci. 123, 3849–3855 (2010).

    Article  CAS  PubMed  Google Scholar 

  5. Schmidt, O., Pfanner, N. & Meisinger, C. Mitochondrial protein import: from proteomics to functional mechanisms. Nature Rev. Mol. Cell Biol. 11, 655–667 (2010).

    Article  CAS  Google Scholar 

  6. Rugarli, E. I. & Langer, T. Mitochondrial quality control: a matter of life and death for neurons. EMBO J. 31, 1336–1349 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Westermann, B. Mitochondrial fusion and fission in cell life and death. Nature Rev. Mol. Cell Biol. 11, 872–884 (2010).

    Article  CAS  Google Scholar 

  8. Youle, R. J. & Narendra, D. P. Mechanisms of mitophagy. Nature Rev. Mol. Cell Biol. 12, 9–14 (2011).

    Article  CAS  Google Scholar 

  9. Nunnari, J. & Suomalainen, A. Mitochondria: in sickness and in health. Cell 148, 1145–1159 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Lin, M. T. & Beal, M. F. Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature 443, 787–795 (2006).

    Article  CAS  PubMed  Google Scholar 

  11. Patti, M. E. & Corvera, S. The role of mitochondria in the pathogenesis of type 2 diabetes. Endocr. Rev. 31, 364–395 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Vafai, S. B. & Mootha, V. K. Mitochondrial disorders as windows into an ancient organelle. Nature 491, 374–383 (2012).

    Article  CAS  PubMed  Google Scholar 

  13. Wallace, D. C. Mitochondrial diseases in man and mouse. Science 283, 1482–1488 (1999).

    Article  CAS  PubMed  Google Scholar 

  14. Shoubridge, E. A. Cytochrome c oxidase deficiency. Am. J. Med. Genet. 106, 46–52 (2001).

    Article  CAS  PubMed  Google Scholar 

  15. Kerstann, K., Brown, M., Vockley, I. & Wallace, D. in The Metabolic and Molecular Bases of Inherited Disease 8th Edn Ch. 105 (eds Valle, P. et al.) (McGraw-Hill, 2001).

    Google Scholar 

  16. Sharpley, M. S. et al. Heteroplasmy of mouse mtDNA is genetically unstable and results in altered behavior and cognition. Cell 151, 333–343 (2012). This paper demonstrates that heteroplasmy for two wild-type mtDNAs is sufficient to alter metabolism and cognitive function in mice.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Jenuth, J. P., Peterson, A. C. & Shoubridge, E. A. Tissue-specific selection for different mtDNA genotypes in heteroplasmic mice. Nature Genet 16, 93–95 (1997). This study reports that mtDNA segregates differently from one tissue to another.

    Article  CAS  PubMed  Google Scholar 

  18. Zeviani, M. & Di Donato, S. Mitochondrial disorders. Brain 127, 2153–2172 (2004).

    Article  PubMed  Google Scholar 

  19. Koene, S. & Smeitink, J. Mitochondrial medicine: entering the era of treatment. J. Intern. Med. 265, 193–209 (2009).

    Article  CAS  PubMed  Google Scholar 

  20. Wallace, D. C., Fan, W. & Procaccio, V. Mitochondrial energetics and therapeutics. Annu. Rev. Pathol. 5, 297–348 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Andreux, P. A. et al. Systems genetics of metabolism: the use of the BXD murine reference panel for multiscalar integration of traits. Cell 150, 1287–1299 (2012). This paper exploits the BXD strains as a potential resource for the complex genetics underlying metabolic phenotypes related to mitochondrial dysfunction.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Kenyon, C. J. The genetics of ageing. Nature 464, 504–512 (2010).

    Article  CAS  PubMed  Google Scholar 

  23. Laplante, M. & Sabatini, D. M. mTOR signaling in growth control and disease. Cell 149, 274–293 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Hardie, D. G., Ross, F. A. & Hawley, S. A. AMPK: a nutrient and energy sensor that maintains energy homeostasis. Nature Rev. Mol. Cell Biol. 13, 251–262 (2012).

    Article  CAS  Google Scholar 

  25. Houtkooper, R. H., Pirinen, E. & Auwerx, J. Sirtuins as regulators of metabolism and healthspan. Nature Rev. Mol. Cell Biol. 13, 225–238 (2012).

    Article  CAS  Google Scholar 

  26. Houtkooper, R. H., Williams, R. W. & Auwerx, J. Metabolic networks of longevity. Cell 142, 9–14 (2010).

    Article  CAS  PubMed  Google Scholar 

  27. Corsetti, G. et al. Morphometric changes induced by amino acid supplementation in skeletal and cardiac muscles of old mice. Am J. Cardiol 101, 26e–34e (2008).

    Article  CAS  PubMed  Google Scholar 

  28. Boudina, S. et al. Contribution of impaired myocardial insulin signaling to mitochondrial dysfunction and oxidative stress in the heart. Circulation 119, 1272–1283 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Hyyti, O. M., Ledee, D., Ning, X. H., Ge, M. & Portman, M. A. Aging impairs myocardial fatty acid and ketone oxidation and modifies cardiac functional and metabolic responses to insulin in mice. Am. J. Physiol. Heart Circ. Physiol. 299, H868–H875 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Houtkooper, R. H. et al. The metabolic footprint of aging in mice. Sci. Rep. 1, 134 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Sinha, R. et al. Assessment of skeletal muscle triglyceride content by 1H nuclear magnetic resonance spectroscopy in lean and obese adolescents: relationships to insulin sensitivity, total body fat, and central adiposity. Diabetes 51, 1022–1027 (2002).

    Article  CAS  PubMed  Google Scholar 

  32. Kotronen, A., Seppala-Lindroos, A., Bergholm, R. & Yki-Jarvinen, H. Tissue specificity of insulin resistance in humans: fat in the liver rather than muscle is associated with features of the metabolic syndrome. Diabetologia 51, 130–138 (2008).

    Article  CAS  PubMed  Google Scholar 

  33. Kelley, D. E., He, J., Menshikova, E. V. & Ritov, V. B. Dysfunction of mitochondria in human skeletal muscle in type 2 diabetes. Diabetes 51, 2944–2950 (2002).

    Article  CAS  PubMed  Google Scholar 

  34. Petersen, K. F., Dufour, S., Befroy, D., Garcia, R. & Shulman, G. I. Impaired mitochondrial activity in the insulin-resistant offspring of patients with type 2 diabetes. N. Engl. J. Med. 350, 664–671 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Hwang, H. et al. Proteomics analysis of human skeletal muscle reveals novel abnormalities in obesity and type 2 diabetes. Diabetes 59, 33–42 (2010).

    Article  CAS  PubMed  Google Scholar 

  36. Mootha, V. K. et al. Integrated analysis of protein composition, tissue diversity, and gene regulation in mouse mitochondria. Cell 115, 629–640 (2003).

    Article  CAS  PubMed  Google Scholar 

  37. Patti, M. E. et al. Coordinated reduction of genes of oxidative metabolism in humans with insulin resistance and diabetes: potential role of PGC1 and NRF1. Proc. Natl Acad. Sci. USA 100, 8466–8471 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Krishnan, J. et al. Dietary obesity-associated Hif1α activation in adipocytes restricts fatty acid oxidation and energy expenditure via suppression of the Sirt2-NAD+ system. Genes Dev. 26, 259–270 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Mustelin, L. et al. Acquired obesity and poor physical fitness impair expression of genes of mitochondrial oxidative phosphorylation in monozygotic twins discordant for obesity. Am. J. Physiol. Endocrinol. Metab. 295, e148–e154 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Pihlajamaki, J. et al. Thyroid hormone-related regulation of gene expression in human fatty liver. J. Clin. Endocrinol. Metab. 94, 3521–3529 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Takamura, T. et al. Obesity upregulates genes involved in oxidative phosphorylation in livers of diabetic patients. Obesity 16, 2601–2609 (2008).

    Article  CAS  PubMed  Google Scholar 

  42. Poussin, C. et al. Oxidative phosphorylation flexibility in the liver of mice resistant to high-fat diet-induced hepatic steatosis. Diabetes 60, 2216–2224 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Auwerx, J. Improving metabolism by increasing energy expenditure. Nature Med. 12, 44–45 (2006).

    Article  CAS  PubMed  Google Scholar 

  44. Meex, R. C. et al. Restoration of muscle mitochondrial function and metabolic flexibility in type 2 diabetes by exercise training is paralleled by increased myocellular fat storage and improved insulin sensitivity. Diabetes 59, 572–579 (2010).

    Article  CAS  PubMed  Google Scholar 

  45. Phielix, E., Meex, R., Moonen-Kornips, E., Hesselink, M. K. & Schrauwen, P. Exercise training increases mitochondrial content and ex vivo mitochondrial function similarly in patients with type 2 diabetes and in control individuals. Diabetologia 53, 1714–1721 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Pospisilik, J. A. et al. Targeted deletion of AIF decreases mitochondrial oxidative phosphorylation and protects from obesity and diabetes. Cell 131, 476–491 (2007).

    Article  CAS  PubMed  Google Scholar 

  47. Vernochet, C. et al. Adipose-specific deletion of TFAM increases mitochondrial oxidation and protects mice against obesity and insulin resistance. Cell Metab. 16, 765–776 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Douglas, P. M. & Dillin, A. Protein homeostasis and aging in neurodegeneration. J. Cell Biol. 190, 719–729 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Valente, E. M. et al. Localization of a novel locus for autosomal recessive early-onset parkinsonism, PARK6, on human chromosome 1p35-p36. Am. J. Hum. Genet. 68, 895–900 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Kitada, T. et al. Mutations in the parkin gene cause autosomal recessive juvenile parkinsonism. Nature 392, 605–608 (1998).

    Article  CAS  PubMed  Google Scholar 

  51. Dagda, R. K. et al. Loss of PINK1 function promotes mitophagy through effects on oxidative stress and mitochondrial fission. J. Biol. Chem. 284, 13843–13855 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Cui, M., Tang, X., Christian, W. V., Yoon, Y. & Tieu, K. Perturbations in mitochondrial dynamics induced by human mutant PINK1 can be rescued by the mitochondrial division inhibitor mdivi-1. J. Biol. Chem. 285, 11740–11752 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Kamp, F. et al. Inhibition of mitochondrial fusion by α-synuclein is rescued by PINK1, Parkin and DJ-1. EMBO J. 29, 3571–3589 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Krebiehl, G. et al. Reduced basal autophagy and impaired mitochondrial dynamics due to loss of Parkinson's disease-associated protein DJ-1. PLoS ONE 5, e9367 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Vives-Bauza, C. et al. PINK1-dependent recruitment of Parkin to mitochondria in mitophagy. Proc. Natl Acad. Sci. USA 107, 378–383 (2010).

    Article  CAS  PubMed  Google Scholar 

  56. Hansson Petersen, C. A. et al. The amyloid β-peptide is imported into mitochondria via the TOM import machinery and localized to mitochondrial cristae. Proc. Natl Acad. Sci. USA 105, 13145–13150 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Barsoum, M. J. et al. Nitric oxide-induced mitochondrial fission is regulated by dynamin-related GTPases in neurons. EMBO J. 25, 3900–3911 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Aleardi, A. M. et al. Gradual alteration of mitochondrial structure and function by β-amyloids: importance of membrane viscosity changes, energy deprivation, reactive oxygen species production, and cytochrome c release. J. Bioenerg. Biomembr. 37, 207–225 (2005).

    Article  CAS  PubMed  Google Scholar 

  59. Crouch, P. J. et al. Copper-dependent inhibition of human cytochrome c oxidase by a dimeric conformer of amyloid-β1–42. J. Neurosci. 25, 672–679 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Area-Gomez, E. et al. Upregulated function of mitochondria-associated ER membranes in Alzheimer disease. EMBO J. 31, 4106–4123 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Coskun, P. et al. A mitochondrial etiology of Alzheimer and Parkinson disease. Biochim. Biophys. Acta 1820, 553–564 (2012).

    Article  CAS  PubMed  Google Scholar 

  62. Fukui, H., Diaz, F., Garcia, S. & Moraes, C. T. Cytochrome c oxidase deficiency in neurons decreases both oxidative stress and amyloid formation in a mouse model of Alzheimer's disease. Proc. Natl Acad. Sci. USA 104, 14163–14168 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Fulda, S., Galluzzi, L. & Kroemer, G. Targeting mitochondria for cancer therapy. Nature Rev. Drug Discov. 9, 447–464 (2010).

    Article  CAS  Google Scholar 

  64. Green, D. R., Galluzzi, L. & Kroemer, G. Mitochondria and the autophagy-inflammation-cell death axis in organismal aging. Science 333, 1109–1112 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Fontana, L., Partridge, L. & Longo, V. D. Extending healthy life span — from yeast to humans. Science 328, 321–326 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Colman, R. J. et al. Caloric restriction delays disease onset and mortality in rhesus monkeys. Science 325, 201–204 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Mattison, J. A. et al. Impact of caloric restriction on health and survival in rhesus monkeys from the NIA study. Nature 489, 318–321 (2012). References 66 and 67 show divergent results of the impact of caloric restriction on health and lifespan in non-human primates.

    Article  CAS  PubMed  Google Scholar 

  68. Imai, S., Armstrong, C. M., Kaeberlein, M. & Guarente, L. Transcriptional silencing and longevity protein Sir2 is an NAD-dependent histone deacetylase. Nature 403, 795–780 (2000).

    Article  CAS  PubMed  Google Scholar 

  69. Matoba, S. et al. p53 regulates mitochondrial respiration. Science 312, 1650–1653 (2006).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Brunet, A. et al. Stress-dependent regulation of FOXO transcription factors by the SIRT1 deacetylase. Science 303, 2011–2015 (2004).

    Article  CAS  PubMed  Google Scholar 

  72. Motta, M. C. et al. Mammalian SIRT1 represses forkhead transcription factors. Cell 116, 551–563 (2004).

    Article  CAS  PubMed  Google Scholar 

  73. van der Horst, A. et al. FOXO4 is acetylated upon peroxide stress and deacetylated by the longevity protein hSir2(SIRT1). J. Biol. Chem. 279, 28873–28879 (2004).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  75. Howitz, K. T. et al. Small molecule activators of sirtuins extend Saccharomyces cerevisiae lifespan. Nature 425, 191–196 (2003).

    Article  CAS  PubMed  Google Scholar 

  76. Beher, D. et al. Resveratrol is not a direct activator of SIRT1 enzyme activity. Chem. Biol. Drug Des. 74, 619–624 (2009).

    Article  CAS  PubMed  Google Scholar 

  77. Canto, C. et al. AMPK regulates energy expenditure by modulating NAD+ metabolism and SIRT1 activity. Nature 458, 1056–1060 (2009). This study defines the importance of the AMPK–SIRT1–PGC1α signaling axis in mitochondrial function.

  78. Canto, C. et al. Interdependence of AMPK and SIRT1 for metabolic adaptation to fasting and exercise in skeletal muscle. Cell Metab. 11, 213–219 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Pacholec, M. et al. SRT1720, SRT2183, SRT1460, and resveratrol are not direct activators of SIRT1. J. Biol. Chem. 285, 8340–8351 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Um, J. H. et al. AMP-activated protein kinase-deficient mice are resistant to the metabolic effects of resveratrol. Diabetes 59, 554–563 (2010).

    Article  CAS  PubMed  Google Scholar 

  81. Baur, J. A. et al. Resveratrol improves health and survival of mice on a high-calorie diet. Nature 444, 337–342 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Lagouge, M. et al. Resveratrol improves mitochondrial function and protects against metabolic disease by activating SIRT1 and PGC-1α. Cell 127, 1109–1122 (2006). References 81 and 82 provide the first evidence of a beneficial impact of resveratrol on mitochondrial function and healthspan in mice.

    Article  CAS  PubMed  Google Scholar 

  83. Timmers, S. et al. Calorie restriction-like effects of 30 days of resveratrol supplementation on energy metabolism and metabolic profile in obese humans. Cell Metab. 14, 612–622 (2011).

    Article  CAS  PubMed  Google Scholar 

  84. Brasnyo, P. et al. Resveratrol improves insulin sensitivity, reduces oxidative stress and activates the Akt pathway in type 2 diabetic patients. Br. J. Nutr. 106, 383–389 (2011).

    Article  CAS  PubMed  Google Scholar 

  85. Poulsen, M. M. et al. High-dose resveratrol supplementation in obese men: an investigator-initiated, randomized, placebo-controlled clinical trial of substrate metabolism, insulin sensitivity, and body composition. Diabetes 62, 1186–1195 (2012).

    Article  CAS  PubMed  Google Scholar 

  86. Yoshino, J. et al. Resveratrol supplementation does not improve metabolic function in nonobese women with normal glucose tolerance. Cell Metab. 16, 658–664 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Bhatt, J. K., Thomas, S. & Nanjan, M. J. Resveratrol supplementation improves glycemic control in type 2 diabetes mellitus. Nutr. Res. 32, 537–541 (2012).

    Article  CAS  PubMed  Google Scholar 

  88. Patel, K. R. et al. Clinical trials of resveratrol. Ann. NY Acad. Sci. 1215, 161–169 (2011).

    Article  CAS  PubMed  Google Scholar 

  89. Milne, J. C. et al. Small molecule activators of SIRT1 as therapeutics for the treatment of type 2 diabetes. Nature 450, 712–716 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Feige, J. N. et al. Specific SIRT1 activation mimics low energy levels and protects against diet-induced metabolic disorders by enhancing fat oxidation. Cell Metab. 8, 347–358 (2008).

    Article  CAS  PubMed  Google Scholar 

  91. Dai, H. et al. SIRT1 activation by small molecules: kinetic and biophysical evidence for direct interaction of enzyme and activator. J. Biol. Chem. 285, 32695–32703 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Baur, J. A., Ungvari, Z., Minor, R. K., Le Couteur, D. G. & de Cabo, R. Are sirtuins viable targets for improving healthspan and lifespan? Nature Rev. Drug Discov. 11, 443–461 (2012).

    Article  CAS  Google Scholar 

  93. Hubbard, B. P. et al. Evidence for a common mechanism of SIRT1 regulation by allosteric activators. Science 339, 1216–1219 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Minor, R. K. et al. SRT1720 improves survival and healthspan of obese mice. Sci. Rep. 1, 70 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Qi, Y. et al. SRT2104, a novel small molecule SIRT1 activator ameliorates insulin resistance and promotes glucose utilization measured under a hyperinsulinemic-euglycemic clamp by enhancing both glycolysis and carbohydrate oxidation in mice fed a high fat diet. Diabetes 60 (Suppl. 1), Abstract 1007-P (2011).

    Google Scholar 

  96. Suri, V. et al. The metabolomic profile of SRT2104, a selective SIRT1 activator in high fat diet fed mice reveals significant improvements in metabolic function in liver, heart and skeletal muscle. Diabetes 60 (Suppl. 1), Abstract 1706-P (2011).

    Google Scholar 

  97. Hoffmann, E. et al. Pharmacokinetics and tolerability of SRT2104, a first-in-class small molecule activator of SIRT1, after single and repeated oral administration in man. Br. J. Clin. Pharmacol. 75, 186–196 (2013).

    Article  CAS  PubMed  Google Scholar 

  98. Bemis, J. E. et al. Discovery of oxazolo[4,5-b]pyridines and related heterocyclic analogs as novel SIRT1 activators. Bioorg. Med. Chem. Lett. 19, 2350–2353 (2009).

    Article  CAS  PubMed  Google Scholar 

  99. Mai, A. et al. Study of 1,4-dihydropyridine structural scaffold: discovery of novel sirtuin activators and inhibitors. J. Med. Chem. 52, 5496–5504 (2009).

    Article  CAS  PubMed  Google Scholar 

  100. Vu, C. B. et al. Discovery of imidazo[1,2-b]thiazole derivatives as novel SIRT1 activators. J. Med. Chem. 52, 1275–1283 (2009).

    Article  CAS  PubMed  Google Scholar 

  101. Houtkooper, R. H., Canto, C., Wanders, R. J. & Auwerx, J. The secret life of NAD+: an old metabolite controlling new metabolic signaling pathways. Endocr. Rev. 31, 194–223 (2010).

    Article  CAS  PubMed  Google Scholar 

  102. Houtkooper, R. H. & Auwerx, J. Exploring the therapeutic space around NAD+. J. Cell Biol. 199, 205–209 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Canto, C. et al. The NAD+ precursor nicotinamide riboside enhances oxidative metabolism and protects against high-fat diet-induced obesity. Cell Metab. 15, 838–847 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Yoshino, J., Mills, K. F., Yoon, M. J. & Imai, S. Nicotinamide mononucleotide, a key NAD+ intermediate, treats the pathophysiology of diet- and age-induced diabetes in mice. Cell Metab. 14, 528–536 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Bai, P. et al. PARP-1 inhibition increases mitochondrial metabolism through SIRT1 activation. Cell Metab. 13, 461–468 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Barbosa, M. T. et al. The enzyme CD38 (a NAD glycohydrolase, EC 3.2.2.5) is necessary for the development of diet-induced obesity. FASEB J. 21, 3629–3639 (2007). References 103–106 (four independent studies) define the importance of NAD+ levels and SIRT activation in mitochondrial homeostasis.

    Article  CAS  PubMed  Google Scholar 

  107. Canto, C. & Auwerx, J. AMP-activated protein kinase and its downstream transcriptional pathways. Cell. Mol. Life Sci. 67, 3407–3423 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Viscomi, C. et al. In vivo correction of COX deficiency by activation of the AMPK/PGC-1α axis. Cell Metab. 14, 80–90 (2011). This paper shows that the AMPK agonist AICAR is beneficial for this form of inherited mitochondrial disease in mice.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Zoncu, R. Efeyan, A. & Sabatini, D. M. mTOR: from growth signal integration to cancer, diabetes and ageing. Nature Rev. Mol. Cell Biol. 12, 21–35 (2011).

    Article  CAS  Google Scholar 

  111. Harrison, D. E. et al. Rapamycin fed late in life extends lifespan in genetically heterogeneous mice. Nature 460, 392–395 (2009). This key report demonstrates the beneficial impact of the mTOR inhibitor rapamycin on mouse lifespan.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Bjedov, I. et al. Mechanisms of life span extension by rapamycin in the fruit fly Drosophila melanogaster. Cell Metab. 11, 35–46 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Robida-Stubbs, S. et al. TOR signaling and rapamycin influence longevity by regulating SKN-1/Nrf and DAF-16/FoxO. Cell Metab. 15, 713–724 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Polak, P. et al. Adipose-specific knockout of raptor results in lean mice with enhanced mitochondrial respiration. Cell Metab. 8, 399–410 (2008).

    Article  CAS  PubMed  Google Scholar 

  115. Bentzinger, C. F. et al. Skeletal muscle-specific ablation of raptor, but not of rictor, causes metabolic changes and results in muscle dystrophy. Cell Metab. 8, 411–424 (2008).

    Article  CAS  PubMed  Google Scholar 

  116. Schoonjans, K., Staels, B. & Auwerx, J. Role of the peroxisome proliferator-activated receptor (PPAR) in mediating the effects of fibrates and fatty acids on gene expression. J. Lipid Res. 37, 907–925 (1996).

    CAS  PubMed  Google Scholar 

  117. Evans, R. M., Barish, G. D. & Wang, Y. X. PPARs and the complex journey to obesity. Nature Med. 10, 355–361 (2004).

    Article  CAS  PubMed  Google Scholar 

  118. Michalik, L. et al. International Union of Pharmacology. LXI. Peroxisome proliferator-activated receptors. Pharmacol. Rev. 58, 726–741 (2006).

    Article  CAS  PubMed  Google Scholar 

  119. Kersten, S. et al. Peroxisome proliferator-activated receptor α mediates the adaptive response to fasting. J. Clin. Invest. 103, 1489–1498 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Leibowitz, M. D. et al. Activation of PPARδ alters lipid metabolism in db/db mice. FEBS Lett. 473, 333–336 (2000).

    Article  CAS  PubMed  Google Scholar 

  121. Tanaka, T. et al. Activation of peroxisome proliferator-activated receptor δ induces fatty acid β-oxidation in skeletal muscle and attenuates metabolic syndrome. Proc. Natl Acad. Sci. USA 100, 15924–15929 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Riserus, U. et al. Activation of peroxisome proliferator-activated receptor (PPAR)δ promotes reversal of multiple metabolic abnormalities, reduces oxidative stress, and increases fatty acid oxidation in moderately obese men. Diabetes 57, 332–339 (2008).

    Article  CAS  PubMed  Google Scholar 

  123. Djouadi, F. et al. Bezafibrate increases very-long-chain acyl-CoA dehydrogenase protein and mRNA expression in deficient fibroblasts and is a potential therapy for fatty acid oxidation disorders. Hum. Mol. Genet. 14, 2695–2703 (2005).

    Article  CAS  PubMed  Google Scholar 

  124. Gobin-Limballe, S. et al. Genetic basis for correction of very-long-chain acyl-coenzyme A dehydrogenase deficiency by bezafibrate in patient fibroblasts: toward a genotype-based therapy. Am. J. Hum. Genet. 81, 1133–1143 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Bonnefont, J. P. et al. Long-term follow-up of bezafibrate treatment in patients with the myopathic form of carnitine palmitoyltransferase 2 deficiency. Clin. Pharmacol. Ther. 88, 101–108 (2010). This study shows that PPARδ activation with bezafibrate has beneficial therapeutic effects in patients with fatty acid oxidation deficiency.

    Article  CAS  PubMed  Google Scholar 

  126. Heikkinen, S., Auwerx, J. & Argmann, C. A. PPARγ in human and mouse physiology. Biochim. Biophys. Acta 1771, 999–1013 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Kung, J. & Henry, R. R. Thiazolidinedione safety. Expert Opin. Drug Saf. 11, 565–579 (2012).

    Article  CAS  PubMed  Google Scholar 

  128. Rocchi, S. et al. A unique PPARγ ligand with potent insulin-sensitizing yet weak adipogenic activity. Mol. Cell 8, 737–747 (2001).

    Article  CAS  PubMed  Google Scholar 

  129. Choi, J. H. et al. Antidiabetic actions of a non-agonist PPARγ ligand blocking Cdk5-mediated phosphorylation. Nature 477, 477–481 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Knouff, C. & Auwerx, J. Peroxisome proliferator-activated receptor-γ calls for activation in moderation: lessons from genetics and pharmacology. Endocr. Rev. 25, 899–918 (2004).

    Article  CAS  PubMed  Google Scholar 

  131. Chae, S. et al. A systems approach for decoding mitochondrial retrograde signaling pathways. Sci. Signal. 6, rs4 (2013).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  133. Dufour, C. R. et al. Genome-wide orchestration of cardiac functions by the orphan nuclear receptors ERRα and γ. Cell Metab. 5, 345–356 (2007).

    Article  CAS  PubMed  Google Scholar 

  134. Schreiber, S. N. et al. The estrogen-related receptor α (ERRα) functions in PPARγ coactivator 1α (PGC-1α)-induced mitochondrial biogenesis. Proc. Natl Acad. Sci. USA 101, 6472–6477 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Alaynick, W. A. et al. ERRγ directs and maintains the transition to oxidative metabolism in the postnatal heart. Cell Metab. 6, 13–24 (2007).

    Article  CAS  PubMed  Google Scholar 

  136. Coward, P., Lee, D., Hull, M. V. & Lehmann, J. M. 4-hydroxytamoxifen binds to and deactivates the estrogen-related receptor γ. Proc. Natl Acad. Sci. USA 98, 8880–8884 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Zuercher, W. J. et al. Identification and structure–activity relationship of phenolic acyl hydrazones as selective agonists for the estrogen-related orphan nuclear receptors ERRβ and ERRγ. J. Med. Chem. 48, 3107–3109 (2005).

    Article  CAS  PubMed  Google Scholar 

  138. Kim, Y., Koh, M., Kim, D. K., Choi, H. S. & Park, S. B. Efficient discovery of selective small molecule agonists of estrogen-related receptor γ using combinatorial approach. J. Comb. Chem. 11, 928–937 (2009).

    Article  CAS  PubMed  Google Scholar 

  139. Willy, P. J. et al. Regulation of PPARγ coactivator 1α (PGC-1α) signaling by an estrogen-related receptor α (ERRα) ligand. Proc. Natl Acad. Sci. USA 101, 8912–8917 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Chisamore, M. J., Cunningham, M. E., Flores, O., Wilkinson, H. A. & Chen, J. D. Characterization of a novel small molecule subtype specific estrogen-related receptor α antagonist in MCF-7 breast cancer cells. PLoS ONE 4, e5624 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Patch, R. J. et al. Identification of diaryl ether-based ligands for estrogen-related receptor α as potential antidiabetic agents. J. Med. Chem. 54, 788–808 (2011).

    Article  CAS  PubMed  Google Scholar 

  142. Kallen, J. et al. Crystal structure of human estrogen-related receptor α in complex with a synthetic inverse agonist reveals its novel molecular mechanism. J. Biol. Chem. 282, 23231–23239 (2007).

    Article  CAS  PubMed  Google Scholar 

  143. Hyatt, S. M. et al. On the intractability of estrogen-related receptor α as a target for activation by small molecules. J. Med. Chem. 50, 6722–6724 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Kim, D. K. et al. Orphan nuclear receptor estrogen-related receptor γ (ERRγ) is key regulator of hepatic gluconeogenesis. J. Biol. Chem. 287, 21628–21639 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Kelly, D. P. & Scarpulla, R. C. Transcriptional regulatory circuits controlling mitochondrial biogenesis and function. Genes Dev. 18, 357–368 (2004).

    Article  CAS  PubMed  Google Scholar 

  146. Collard, J. & Khorkova Sherman, O. Treatment of nuclear respiratory factor 1 (NRF1) related diseases by inhibition of natural antisense transcript to NRF1. US Patent 20120264812 A1 (2010).

  147. Iyer, S. et al. Recombinant mitochondrial transcription factor A with N-terminal mitochondrial transduction domain increases respiration and mitochondrial gene expression. Mitochondrion 9, 196–203 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Thomas, R. R. et al. RhTFAM treatment stimulates mitochondrial oxidative metabolism and improves memory in aged mice. Aging 4, 620–635 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Iyer, S. et al. Mitochondrial gene therapy improves respiration, biogenesis, and transcription in G11778A Leber's hereditary optic neuropathy and T8993G Leigh's syndrome cells. Hum. Gene Ther. 23, 647–657 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Ylikallio, E., Tyynismaa, H., Tsutsui, H., Ide, T. & Suomalainen, A. High mitochondrial DNA copy number has detrimental effects in mice. Hum. Mol. Genet. 19, 2695–2705 (2010).

    Article  CAS  PubMed  Google Scholar 

  151. Picard, F. et al. SRC-1 and TIF2 control energy balance between white and brown adipose tissues. Cell 111, 931–941 (2002).

    Article  CAS  PubMed  Google Scholar 

  152. Coste, A. et al. The genetic ablation of SRC-3 protects against obesity and improves insulin sensitivity by reducing the acetylation of PGC-1α. Proc. Natl Acad. Sci. USA 105, 17187–17192 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Coste, A. et al. Absence of the steroid receptor coactivator-3 induces B-cell lymphoma. EMBO J. 25, 2453–2464 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Choi, J. H. et al. Anti-diabetic drugs inhibit obesity-linked phosphorylation of PPARγ by Cdk5. Nature 466, 451–456 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Ye, L. et al. TRPV4 is a regulator of adipose oxidative metabolism, inflammation, and energy homeostasis. Cell 151, 96–110 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Bostrom, P. et al. A PGC1-α-dependent myokine that drives brown-fat-like development of white fat and thermogenesis. Nature 481, 463–468 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Watanabe, M. et al. Bile acids induce energy expenditure by promoting intracellular thyroid hormone activation. Nature 439, 484–489 (2006).

    Article  CAS  PubMed  Google Scholar 

  158. Sengupta, S., Peterson, T. R., Laplante, M., Oh, S. & Sabatini, D. M. mTORC1 controls fasting-induced ketogenesis and its modulation by ageing. Nature 468, 1100–1104 (2010).

    Article  CAS  PubMed  Google Scholar 

  159. Li, P. et al. Adipocyte NCoR knockout decreases PPARγ phosphorylation and enhances PPARγ activity and insulin sensitivity. Cell 147, 815–826 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Iezzi, S. et al. Deacetylase inhibitors increase muscle cell size by promoting myoblast recruitment and fusion through induction of follistatin. Dev. Cell 6, 673–684 (2004).

    Article  CAS  PubMed  Google Scholar 

  161. Fajas, L. et al. The retinoblastoma–histone deacetylase 3 complex inhibits PPARγ and adipocyte differentiation. Dev. Cell 3, 903–910 (2002).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  163. Debevec, D. et al. Receptor interacting protein 140 regulates expression of uncoupling protein 1 in adipocytes through specific peroxisome proliferator activated receptor isoforms and estrogen-related receptor α. Mol. Endocrinol. 21, 1581–1592 (2007).

    Article  CAS  PubMed  Google Scholar 

  164. Seth, A. et al. The transcriptional corepressor RIP140 regulates oxidative metabolism in skeletal muscle. Cell Metab. 6, 236–245 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Leonardsson, G. et al. Nuclear receptor corepressor RIP140 regulates fat accumulation. Proc. Natl Acad. Sci. USA 101, 8437–8442 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Powelka, A. M. et al. Suppression of oxidative metabolism and mitochondrial biogenesis by the transcriptional corepressor RIP140 in mouse adipocytes. J. Clin. Invest. 116, 125–136 (2006).

    Article  CAS  PubMed  Google Scholar 

  167. Blanchet, E. et al. E2F transcription factor-1 regulates oxidative metabolism. Nature Cell Biol. 13, 1146–1152 (2011).

    Article  CAS  PubMed  Google Scholar 

  168. Scime, A. et al. Rb and p107 regulate preadipocyte differentiation into white versus brown fat through repression of PGC-1α. Cell Metab. 2, 283–295 (2005).

    Article  CAS  PubMed  Google Scholar 

  169. Dali-Youcef, N. et al. Adipose tissue-specific inactivation of the retinoblastoma protein protects against diabesity because of increased energy expenditure. Proc. Natl Acad. Sci. USA 104, 10703–10708 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Calo, E. et al. Rb regulates fate choice and lineage commitment in vivo. Nature 466, 1110–1114 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  171. White, R. et al. The nuclear receptor co-repressor nrip1 (RIP140) is essential for female fertility. Nature Med. 6, 1368–1374 (2000).

    Article  CAS  PubMed  Google Scholar 

  172. Brand, M. D. The sites and topology of mitochondrial superoxide production. Exp. Gerontol. 45, 466–472 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Sena, L. A. & Chandel, N. S. Physiological roles of mitochondrial reactive oxygen species. Mol. Cell 48, 158–167 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Brand, M. D. Uncoupling to survive? The role of mitochondrial inefficiency in ageing. Exp. Gerontol. 35, 811–820 (2000).

    Article  CAS  PubMed  Google Scholar 

  175. Chen, H. et al. Mitofusins Mfn1 and Mfn2 coordinately regulate mitochondrial fusion and are essential for embryonic development. J. Cell Biol. 160, 189–200 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Chan, D. C. Mitochondria: dynamic organelles in disease, aging, and development. Cell 125, 1241–1252 (2006).

    Article  CAS  PubMed  Google Scholar 

  177. Alexander, C. et al. OPA1, encoding a dynamin-related GTPase, is mutated in autosomal dominant optic atrophy linked to chromosome 3q28. Nature Genet. 26, 211–215 (2000).

    Article  CAS  PubMed  Google Scholar 

  178. Zuchner, S. et al. Mutations in the mitochondrial GTPase mitofusin 2 cause Charcot–Marie–Tooth neuropathy type 2A. Nature Genet. 36, 449–451 (2004).

    Article  CAS  PubMed  Google Scholar 

  179. Waterham, H. R. et al. A lethal defect of mitochondrial and peroxisomal fission. N. Engl. J. Med. 356, 1736–1741 (2007).

    Article  CAS  PubMed  Google Scholar 

  180. Ishihara, N. et al. Mitochondrial fission factor Drp1 is essential for embryonic development and synapse formation in mice. Nature Cell Biol. 11, 958–966 (2009).

    Article  CAS  PubMed  Google Scholar 

  181. Chen, H., Chomyn, A. & Chan, D. C. Disruption of fusion results in mitochondrial heterogeneity and dysfunction. J. Biol. Chem. 280, 26185–26192 (2005).

    Article  CAS  PubMed  Google Scholar 

  182. Parone, P. A. et al. Preventing mitochondrial fission impairs mitochondrial function and leads to loss of mitochondrial DNA. PLoS ONE 3, e3257 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Soriano, F. X. et al. Evidence for a mitochondrial regulatory pathway defined by peroxisome proliferator-activated receptor-γ coactivator-1α, estrogen-related receptor-α, and mitofusin 2. Diabetes 55, 1783–1791 (2006).

    Article  CAS  PubMed  Google Scholar 

  184. Cartoni, R. et al. Mitofusins 1/2 and ERRα expression are increased in human skeletal muscle after physical exercise. J. Physiol. 567, 349–358 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Wang, D. et al. A small molecule promotes mitochondrial fusion in mammalian cells. Angew. Chem. Int. Ed Engl. 51, 9302–9305 (2012). This paper reports the identification of a small molecule that stimulates mitochondrial fusion in mammalian cells.

    Article  CAS  PubMed  Google Scholar 

  186. Cassidy-Stone, A. et al. Chemical inhibition of the mitochondrial division dynamin reveals its role in Bax/Bak-dependent mitochondrial outer membrane permeabilization. Dev. Cell 14, 193–204 (2008). This study identifies a small molecule that inhibits mitochondrial fission in yeast.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Lackner, L. L. & Nunnari, J. Small molecule inhibitors of mitochondrial division: tools that translate basic biological research into medicine. Chem. Biol. 17, 578–583 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Nakatogawa, H., Suzuki, K., Kamada, Y. & Ohsumi, Y. Dynamics and diversity in autophagy mechanisms: lessons from yeast. Nature Rev. Mol. Cell Biol. 10, 458–467 (2009).

    Article  CAS  Google Scholar 

  189. Yang, Z. & Klionsky, D. J. Eaten alive: a history of macroautophagy. Nature Cell Biol. 12, 814–822 (2010).

    Article  CAS  PubMed  Google Scholar 

  190. Sandoval, H. et al. Essential role for Nix in autophagic maturation of erythroid cells. Nature 454, 232–235 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Narendra, D. P. et al. PINK1 is selectively stabilized on impaired mitochondria to activate Parkin. PLoS Biol. 8, e1000298 (2010). This key paper demonstrates the recruitment of PINK1 as a stimulator of mitophagy.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Geisler, S. et al. PINK1/Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nature Cell Biol. 12, 119–131 (2010).

    Article  CAS  PubMed  Google Scholar 

  193. Lee, J. Y., Nagano, Y., Taylor, J. P., Lim, K. L. & Yao, T. P. Disease-causing mutations in parkin impair mitochondrial ubiquitination, aggregation, and HDAC6-dependent mitophagy. J. Cell Biol. 189, 671–679 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Matsuda, N. et al. PINK1 stabilized by mitochondrial depolarization recruits Parkin to damaged mitochondria and activates latent Parkin for mitophagy. J. Cell Biol. 189, 211–221 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Sterky, F. H., Lee, S., Wibom, R., Olson, L. & Larsson, N. G. Impaired mitochondrial transport and Parkin-independent degeneration of respiratory chain-deficient dopamine neurons in vivo. Proc. Natl Acad. Sci. USA 108, 12937–12942 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  196. Gegg, M. E. et al. Mitofusin 1 and mitofusin 2 are ubiquitinated in a PINK1/parkin-dependent manner upon induction of mitophagy. Hum. Mol. Genet. 19, 4861–4870 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Tanaka, A. et al. Proteasome and p97 mediate mitophagy and degradation of mitofusins induced by Parkin. J. Cell Biol. 191, 1367–1380 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Hartl, F. U., Bracher, A. & Hayer-Hartl, M. Molecular chaperones in protein folding and proteostasis. Nature 475, 324–332 (2011).

    Article  CAS  PubMed  Google Scholar 

  199. Nargund, A. M., Pellegrino, M. W., Fiorese, C. J., Baker, B. M. & Haynes, C. M. Mitochondrial import efficiency of ATFS-1 regulates mitochondrial UPR activation. Science 337, 587–590 (2012). This report describes another layer of transcriptional control of the mitochondrial UPR by regulating the subcellular distribution of ATFS-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. Benedetti, C., Haynes, C. M., Yang, Y., Harding, H. P. & Ron, D. Ubiquitin-like protein 5 positively regulates chaperone gene expression in the mitochondrial unfolded protein response. Genetics 174, 229–239 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  201. Durieux, J., Wolff, S. & Dillin, A. The cell-non-autonomous nature of electron transport chain-mediated longevity. Cell 144, 79–91 (2011). This study demonstrates that lifespan extension in C. elegans following mitochondrial dysfunction in the neurons requires the mitochondrial UPR in distal tissues. It also defines the concept of a mitokine.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Tyynismaa, H. et al. Mitochondrial myopathy induces a starvation-like response. Hum. Mol. Genet. 19, 3948–3958 (2010).

    Article  CAS  PubMed  Google Scholar 

  203. Suomalainen, A. et al. FGF-21 as a biomarker for muscle-manifesting mitochondrial respiratory chain deficiencies: a diagnostic study. Lancet Neurol. 10, 806–818 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. Kim, K. H. et al. Autophagy deficiency leads to protection from obesity and insulin resistance by inducing Fgf21 as a mitokine. Nature Med. 19, 83–92 (2012). References 203 and 204 suggest that FGF21 can function as a mitokine.

    Article  CAS  PubMed  Google Scholar 

  205. Lee, C., Yen, K. & Cohen, P. Humanin: a harbinger of mitochondrial-derived peptides? Trends Endocrinol. Metab. 8 Feb 2013 (10.1016/j.tem.2013.01.005).

  206. Martinus, R. D. et al. Selective induction of mitochondrial chaperones in response to loss of the mitochondrial genome. Eur. J. Biochem. 240, 98–103 (1996).

    Article  CAS  PubMed  Google Scholar 

  207. Houtkooper, R. H. et al. Mitonuclear protein imbalance as a conserved longevity mechanism. Nature (in the press). This recently published paper shows that the imbalance between mitochondrial and nuclear proteins triggers the mitochondrial UPR, which is an overarching mechanism that enhances lifespan.

  208. Fontaine, B. et al. A new locus for autosomal dominant pure spastic paraplegia, on chromosome 2q24-q34. Am. J. Hum. Genet. 66, 702–707 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  209. Casari, G. et al. Spastic paraplegia and OXPHOS impairment caused by mutations in paraplegin, a nuclear-encoded mitochondrial metalloprotease. Cell 93, 973–983 (1998).

    Article  CAS  PubMed  Google Scholar 

  210. Nolden, M. et al. The m-AAA protease defective in hereditary spastic paraplegia controls ribosome assembly in mitochondria. Cell 123, 277–289 (2005).

    Article  CAS  PubMed  Google Scholar 

  211. Ventura, N. et al. p53/CEP-1 increases or decreases lifespan, depending on level of mitochondrial bioenergetic stress. Aging Cell 8, 380–393 (2009).

    Article  CAS  PubMed  Google Scholar 

  212. Guillon, B. et al. Frataxin deficiency causes upregulation of mitochondrial Lon and ClpP proteases and severe loss of mitochondrial Fe-S proteins. FEBS J. 276, 1036–1047 (2009).

    Article  CAS  PubMed  Google Scholar 

  213. Rotig, A. et al. Aconitase and mitochondrial iron–sulphur protein deficiency in Friedreich ataxia. Nature Genet. 17, 215–217 (1997).

    Article  CAS  PubMed  Google Scholar 

  214. Bieganowski, P. & Brenner, C. Discoveries of nicotinamide riboside as a nutrient and conserved NRK genes establish a Preiss-Handler independent route to NAD+ in fungi and humans. Cell 117, 495–502 (2004).

    Article  CAS  PubMed  Google Scholar 

  215. Sato, H. et al. Anti-hyperglycemic activity of a TGR5 agonist isolated from Olea europaea. Biochem. Biophys. Res. Commun. 362, 793–798 (2007).

    Article  CAS  PubMed  Google Scholar 

  216. Sato, H. et al. Novel potent and selective bile acid derivatives as TGR5 agonists: biological screening, structure–activity relationships, and molecular modeling studies. J. Med. Chem. 51, 1831–1841 (2008).

    Article  CAS  PubMed  Google Scholar 

  217. Evans, K. A. et al. Discovery of 3-aryl-4-isoxazolecarboxamides as TGR5 receptor agonists. J. Med. Chem. 52, 7962–7965 (2009).

    Article  CAS  PubMed  Google Scholar 

  218. Pellicciari, R. et al. Discovery of 6α-ethyl-23(S)-methylcholic acid (S-EMCA, INT-777) as a potent and selective agonist for the TGR5 receptor, a novel target for diabesity. J. Med. Chem. 52, 7958–7961 (2009).

    Article  CAS  PubMed  Google Scholar 

  219. Genet, C. et al. Structure–activity relationship study of betulinic acid, a novel and selective TGR5 agonist, and its synthetic derivatives: potential impact in diabetes. J. Med. Chem. 53, 178–190 (2010).

    Article  CAS  PubMed  Google Scholar 

  220. Herbert, M. R. et al. Synthesis and SAR of 2-aryl-3-aminomethylquinolines as agonists of the bile acid receptor TGR5. Bioorg. Med. Chem. Lett. 20, 5718–5721 (2010).

    Article  CAS  PubMed  Google Scholar 

  221. Londregan, A. T. et al. Discovery of 5-phenoxy-1,3-dimethyl-1H-pyrazole-4-carboxamides as potent agonists of TGR5 via sequential combinatorial libraries. Bioorg. Med. Chem. Lett. 23, 1407–1411 (2013).

    Article  CAS  PubMed  Google Scholar 

  222. Moreno, D., Knecht, E., Viollet, B. & Sanz, P. A769662, a novel activator of AMP-activated protein kinase, inhibits non-proteolytic components of the 26S proteasome by an AMPK-independent mechanism. FEBS Lett. 582, 2650–2654 (2008).

    Article  CAS  PubMed  Google Scholar 

  223. Hawley, S. A. et al. The ancient drug salicylate directly activates AMP-activated protein kinase. Science 336, 918–922 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  224. Cairns, R. A., Harris, I. S. & Mak, T. W. Regulation of cancer cell metabolism. Nature Rev. Cancer 11, 85–95 (2011).

    Article  CAS  Google Scholar 

  225. Kitami, T. et al. A chemical screen probing the relationship between mitochondrial content and cell size. PLoS ONE 7, e33755 (2012). This study reports the set-up of an image-based phenotypic screen for the identification of compounds that increase mitochondrial content per cell using a primary cell line.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  226. Beeson, C. C., Beeson, G. C. & Schnellmann, R. G. A high-throughput respirometric assay for mitochondrial biogenesis and toxicity. Anal. Biochem. 404, 75–81 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Kikkawa, R., Yamamoto, T., Fukushima, T., Yamada, H. & Horii, I. Investigation of a hepatotoxicity screening system in primary cell cultures — “what biomarkers would need to be addressed to estimate toxicity in conventional and new approaches?”. J. Toxicol. Sci. 30, 61–72 (2005).

    Article  CAS  PubMed  Google Scholar 

  228. Li, A. P. Screening for human ADME/Tox drug properties in drug discovery. Drug Discov. Today 6, 357–366 (2001).

    Article  CAS  PubMed  Google Scholar 

  229. Couplan, E. et al. A yeast-based assay identifies drugs active against human mitochondrial disorders. Proc. Natl Acad. Sci. USA 108, 11989–11994 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  230. Burns, A. R. et al. High-throughput screening of small molecules for bioactivity and target identification in Caenorhabditis elegans. Nature Protoc. 1, 1906–1914 (2006). References 229 and 230 are examples of phenotypic screens for mitochondrial function performed in C. elegans and yeast.

    Article  CAS  Google Scholar 

  231. Segalat, L. Invertebrate animal models of diseases as screening tools in drug discovery. ACS Chem. Biol. 2, 231–236 (2007).

    Article  CAS  PubMed  Google Scholar 

  232. Gohil, V. M. et al. Nutrient-sensitized screening for drugs that shift energy metabolism from mitochondrial respiration to glycolysis. Nature Biotech. 28, 249–255 (2010).

    Article  CAS  Google Scholar 

  233. Wagner, B. K. et al. Large-scale chemical dissection of mitochondrial function. Nature Biotech. 26, 343–351 (2008). References 232 and 233 describe two examplary phenotypic screens for mitochondrial function in mammalian cells.

    Article  CAS  Google Scholar 

  234. Tsang, W. Y. & Lemire, B. D. Mitochondrial genome content is regulated during nematode development. Biochem. Biophys. Res. Commun. 291, 8–16 (2002).

    Article  CAS  PubMed  Google Scholar 

  235. Detmer, S. A. & Chan, D. C. Functions and dysfunctions of mitochondrial dynamics. Nature Rev. Mol. Cell Biol. 8, 870–879 (2007).

    Article  CAS  Google Scholar 

  236. Wellcome Trust Case Control Consortium. Genome-wide association study of 14,000 cases of seven common diseases and 3,000 shared controls. Nature 447, 661–678 (2007).

  237. Frayling, T. M. Genome-wide association studies provide new insights into type 2 diabetes aetiology. Nature Rev. Genet. 8, 657–662 (2007).

    Article  CAS  PubMed  Google Scholar 

  238. Zeggini, E. et al. Meta-analysis of genome-wide association data and large-scale replication identifies additional susceptibility loci for type 2 diabetes. Nature Genet. 40, 638–645 (2008).

    Article  CAS  PubMed  Google Scholar 

  239. Mair, W. & Dillin, A. Aging and survival: the genetics of life span extension by dietary restriction. Annu. Rev. Biochem. 77, 727–754 (2008).

    Article  CAS  PubMed  Google Scholar 

  240. Handschin, C. & Spiegelman, B. M. Peroxisome proliferator-activated receptor γ coactivator 1 coactivators, energy homeostasis, and metabolism. Endocr. Rev. 27, 728–735 (2006).

    Article  CAS  PubMed  Google Scholar 

  241. Fernandez-Marcos, P. J. & Auwerx, J. Regulation of PGC-1α, a nodal regulator of mitochondrial biogenesis. Am J. Clin. Nutr. 93, 884S–890S (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  242. Canto, C. & Auwerx, J. PGC-1α, SIRT1 and AMPK, an energy sensing network that controls energy expenditure. Curr. Opin. Lipidol. 20, 98–105 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  243. Jeninga, E. H., Schoonjans, K. & Auwerx, J. Reversible acetylation of PGC-1: connecting energy sensors and effectors to guarantee metabolic flexibility. Oncogene 29, 4617–4624 (2010).

    Article  CAS  PubMed  Google Scholar 

  244. Feige, J. N., Gelman, L., Michalik, L., Desvergne, B. & Wahli, W. From molecular action to physiological outputs: peroxisome proliferator-activated receptors are nuclear receptors at the crossroads of key cellular functions. Prog. Lipid Res. 45, 120–159 (2006).

    Article  CAS  PubMed  Google Scholar 

  245. Feige, J. N. & Auwerx, J. Transcriptional coregulators in the control of energy homeostasis. Trends Cell Biol. 17, 292–301 (2007).

    Article  CAS  PubMed  Google Scholar 

  246. Scarpulla, R. C., Vega, R. B. & Kelly, D. P. Transcriptional integration of mitochondrial biogenesis. Trends Endocrinol. Metab. 23, 459–466 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  247. Kalyanaraman, B. et al. Measuring reactive oxygen and nitrogen species with fluorescent probes: challenges and limitations. Free Radic. Biol. Med. 52, 1–6 (2012).

    Article  CAS  PubMed  Google Scholar 

  248. Meyer, A. J. & Dick, T. P. Fluorescent protein-based redox probes. Antioxid. Redox Signal. 13, 621–650 (2010).

    Article  CAS  PubMed  Google Scholar 

  249. Greer, E. L. et al. An AMPK–FOXO pathway mediates longevity induced by a novel method of dietary restriction in C. elegans. Curr. Biol. 17, 1646–1656 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank R. Wanders for critically reading the manuscript. R.H.H. is financially supported by an AMC Postdoctoral fellowship and a ZonMw-VENI grant (number 91613050). J.A. is the Nestlé Chair in Energy Metabolism and work in his laboratory is supported by the École polytechnique fédérale de Lausanne (EPFL), the EU Ideas program (ERC-2008-AdG-231138), the US National Institutes of Health (grants 1R01HL 106511-01A1 and R01AG043930), the Velux Stiftung Research Grant Program and the Swiss National Science Foundation (grants 31003A-124713 and CRSII3-136201).

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Riekelt H. Houtkooper or Johan Auwerx.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

FURTHER INFORMATION

Johan Auwerx's homepage

PowerPoint slides

Glossary

Mitochondrial DNA

(mtDNA). A 16.5 kb circular DNA sequence carried within mitochondria, composed of a light and heavy strand. Both strands contain 37 genes, including 13 that encode protein subunits of the oxidative phosphorylation complexes, whereas the remaining encode ribosomal RNA and tRNA molecules that are essential for the transcription and synthesis of mitochondrially encoded proteins.

Oxidative phosphorylation

Enzymatic phosphorylation of ADP to ATP, which is coupled to electron transfer from a substrate to molecular oxygen in the electron transport chain.

Heteroplasmy

A mixture of different forms of mitochondrial DNA within a single cell.

Quantitative trait loci

(QTLs). Genetic loci contributing quantitatively to a trait or phenotype. A QTL partly explains the genetic contribution to a given phenotype.

Proteostasis

Homeostasis of the protein folding landscape.

mtDNA haplogroup

A group sharing the same single nucleotide polymorphism on mitochondrial DNA (mtDNA). Some haplogroups have been suggested to have a greater likelihood of predisposing individuals to Alzheimer's disease.

Caloric restriction diet

A diet that involves the consumption of 20–50% less calories than normal without causing malnutrition: that is, with maintenance of proper vitamin and mineral intake.

NAD+

The oxidized version of NAD; serves as a co-factor in oxidation–reduction reactions. NAD+ also acts as an obligatory co-substrate for sirtuin-mediated deacylation reactions.

Inverse agonists

Compounds that bind to the same receptor binding site as a prototypical agonist for that given receptor. Through this binding, inverse agonists reverse the basal or constitutive activity of the receptor. Oestrogen-related receptor-α is an example of a constitutively active nuclear receptor for which inverse agonists exist.

Myokine

A signalling molecule that is produced by the muscle and has cytokine-like properties.

Mitokine

A signalling molecule that is produced by mitochondria in a given tissue and affects mitochondrial function in a distinct tissue.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Andreux, P., Houtkooper, R. & Auwerx, J. Pharmacological approaches to restore mitochondrial function. Nat Rev Drug Discov 12, 465–483 (2013). https://doi.org/10.1038/nrd4023

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrd4023

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