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.

  • Letter
  • Published:

The metabolite α-ketoglutarate extends lifespan by inhibiting ATP synthase and TOR

Abstract

Metabolism and ageing are intimately linked. Compared with ad libitum feeding, dietary restriction consistently extends lifespan and delays age-related diseases in evolutionarily diverse organisms1,2. Similar conditions of nutrient limitation and genetic or pharmacological perturbations of nutrient or energy metabolism also have longevity benefits3,4. Recently, several metabolites have been identified that modulate ageing5,6; however, the molecular mechanisms underlying this are largely undefined. Here we show that α-ketoglutarate (α-KG), a tricarboxylic acid cycle intermediate, extends the lifespan of adult Caenorhabditis elegans. ATP synthase subunit β is identified as a novel binding protein of α-KG using a small-molecule target identification strategy termed drug affinity responsive target stability (DARTS)7. The ATP synthase, also known as complex V of the mitochondrial electron transport chain, is the main cellular energy-generating machinery and is highly conserved throughout evolution8,9. Although complete loss of mitochondrial function is detrimental, partial suppression of the electron transport chain has been shown to extend C. elegans lifespan10,11,12,13. We show that α-KG inhibits ATP synthase and, similar to ATP synthase knockdown, inhibition by α-KG leads to reduced ATP content, decreased oxygen consumption, and increased autophagy in both C. elegans and mammalian cells. We provide evidence that the lifespan increase by α-KG requires ATP synthase subunit β and is dependent on target of rapamycin (TOR) downstream. Endogenous α-KG levels are increased on starvation and α-KG does not extend the lifespan of dietary-restricted animals, indicating that α-KG is a key metabolite that mediates longevity by dietary restriction. Our analyses uncover new molecular links between a common metabolite, a universal cellular energy generator and dietary restriction in the regulation of organismal lifespan, thus suggesting new strategies for the prevention and treatment of ageing and age-related diseases.

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

Access options

Buy this article

Purchase on Springer Link

Instant access to full article PDF

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

Figure 1: α-KG extends the adult lifespan of C. elegans.
Figure 2: α-KG binds and inhibits ATP synthase.
Figure 3: α-KG longevity is mediated through ATP synthase and the dietary restriction/TOR axis.
Figure 4: Inhibition of ATP synthase by α-KG causes a conserved decrease in TOR pathway activity.

Similar content being viewed by others

References

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

    CAS  PubMed  PubMed Central  ADS  Google Scholar 

  2. 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)

    CAS  PubMed  ADS  Google Scholar 

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

    CAS  PubMed  ADS  Google Scholar 

  4. Harrison, D. E. et al. Rapamycin fed late in life extends lifespan in genetically heterogeneous mice. Nature 460, 392–395 (2009)

    CAS  PubMed  PubMed Central  ADS  Google Scholar 

  5. Williams, D. S., Cash, A., Hamadani, L. & Diemer, T. Oxaloacetate supplementation increases lifespan in Caenorhabditis elegans through an AMPK/FOXO-dependent pathway. Aging Cell 8, 765–768 (2009)

    CAS  PubMed  Google Scholar 

  6. Lucanic, M. et al. N-acylethanolamine signalling mediates the effect of diet on lifespan in Caenorhabditis elegans. Nature 473, 226–229 (2011)

    CAS  PubMed  PubMed Central  ADS  Google Scholar 

  7. Lomenick, B. et al. Target identification using drug affinity responsive target stability (DARTS). Proc. Natl Acad. Sci. USA 106, 21984–21989 (2009)

    CAS  PubMed  ADS  PubMed Central  Google Scholar 

  8. Abrahams, J. P., Leslie, A. G., Lutter, R. & Walker, J. E. Structure at 2.8 Å resolution of F1-ATPase from bovine heart mitochondria. Nature 370, 621–628 (1994)

    CAS  PubMed  ADS  Google Scholar 

  9. Boyer, P. D. The ATP synthase—a splendid molecular machine. Annu. Rev. Biochem. 66, 717–749 (1997)

    CAS  PubMed  Google Scholar 

  10. Tsang, W. Y., Sayles, L. C., Grad, L. I., Pilgrim, D. B. & Lemire, B. D. Mitochondrial respiratory chain deficiency in Caenorhabditis elegans results in developmental arrest and increased life span. J. Biol. Chem. 276, 32240–32246 (2001)

    CAS  PubMed  Google Scholar 

  11. Dillin, A. et al. Rates of behavior and aging specified by mitochondrial function during development. Science 298, 2398–2401 (2002)

    CAS  PubMed  ADS  Google Scholar 

  12. Lee, S. S. et al. A systematic RNAi screen identifies a critical role for mitochondria in C. elegans longevity. Nature Genet. 33, 40–48 (2002)

    PubMed  Google Scholar 

  13. Curran, S. P. & Ruvkun, G. Lifespan regulation by evolutionarily conserved genes essential for viability. PLoS Genet. 3, e56 (2007)

    PubMed  PubMed Central  Google Scholar 

  14. Gems, D. & Riddle, D. L. Genetic, behavioral and environmental determinants of male longevity in Caenorhabditis elegans. Genetics 154, 1597–1610 (2000)

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Brand, M. D. & Nicholls, D. G. Assessing mitochondrial dysfunction in cells. Biochem. J. 435, 297–312 (2011)

    CAS  PubMed  Google Scholar 

  16. Lakowski, B. & Hekimi, S. The genetics of caloric restriction in Caenorhabditis elegans. Proc. Natl Acad. Sci. USA 95, 13091–13096 (1998)

    CAS  PubMed  ADS  PubMed Central  Google Scholar 

  17. Hansen, M. et al. Lifespan extension by conditions that inhibit translation in Caenorhabditis elegans. Aging Cell 6, 95–110 (2007)

    CAS  PubMed  Google Scholar 

  18. Stanfel, M. N., Shamieh, L. S., Kaeberlein, M. & Kennedy, B. K. The TOR pathway comes of age. Biochim. Biophys. Acta 1790, 1067–1074 (2009)

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Hardie, D. G., Scott, J. W., Pan, D. A. & Hudson, E. R. Management of cellular energy by the AMP-activated protein kinase system. FEBS Lett. 546, 113–120 (2003)

    CAS  PubMed  Google Scholar 

  20. Greer, E. L. & Brunet, A. Different dietary restriction regimens extend lifespan by both independent and overlapping genetic pathways in C. elegans. Aging Cell 8, 113–127 (2009)

    CAS  PubMed  Google Scholar 

  21. Sheaffer, K. L., Updike, D. L. & Mango, S. E. The target of rapamycin pathway antagonizes pha-4/FoxA to control development and aging. Curr. Biol. 18, 1355–1364 (2008)

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Panowski, S. H., Wolff, S., Aguilaniu, H., Durieux, J. & Dillin, A. PHA-4/Foxa mediates diet-restriction-induced longevity of C. elegans. Nature 447, 550–555 (2007)

    CAS  PubMed  ADS  Google Scholar 

  23. Wullschleger, S., Loewith, R. & Hall, M. N. TOR signaling in growth and metabolism. Cell 124, 471–484 (2006)

    CAS  PubMed  Google Scholar 

  24. Meléndez, A. et al. Autophagy genes are essential for dauer development and life-span extension in C. elegans. Science 301, 1387–1391 (2003)

    PubMed  ADS  Google Scholar 

  25. Loenarz, C. & Schofield, C. J. Expanding chemical biology of 2-oxoglutarate oxygenases. Nature Chem. Biol. 4, 152–156 (2008)

    CAS  Google Scholar 

  26. Epstein, A. C. et al. C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation. Cell 107, 43–54 (2001)

    CAS  PubMed  Google Scholar 

  27. Zhang, Y., Shao, Z., Zhai, Z., Shen, C. & Powell-Coffman, J. A. The HIF-1 hypoxia-inducible factor modulates lifespan in C. elegans. PLoS ONE 4, e6348 (2009)

    PubMed  PubMed Central  ADS  Google Scholar 

  28. Brauer, M. J. et al. Conservation of the metabolomic response to starvation across two divergent microbes. Proc. Natl Acad. Sci. USA 103, 19302–19307 (2006)

    CAS  PubMed  ADS  PubMed Central  Google Scholar 

  29. Kaminsky, Y. G., Kosenko, E. A. & Kondrashova, M. N. Metabolites of citric acid cycle, carbohydrate and phosphorus metabolism, and related reactions, redox and phosphorylating states of hepatic tissue, liver mitochondria and cytosol of the pigeon, under normal feeding and natural nocturnal fasting conditions. Comp. Biochem. Physiol. B 73, 957–963 (1982)

    CAS  PubMed  Google Scholar 

  30. Brugnara, L. et al. Metabolomics approach for analyzing the effects of exercise in subjects with type 1 diabetes mellitus. PLoS ONE 7, e40600 (2012)

    CAS  PubMed  PubMed Central  ADS  Google Scholar 

  31. Brenner, S. The genetics of Caenorhabditis elegans. Genetics 77, 71–94 (1974)

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Timmons, L. & Fire, A. Specific interference by ingested dsRNA. Nature 395, 854 (1998)

    CAS  PubMed  ADS  Google Scholar 

  33. Long, X. et al. TOR deficiency in C. elegans causes developmental arrest and intestinal atrophy by inhibition of mRNA translation. Curr. Biol. 12, 1448–1461 (2002)

    CAS  PubMed  Google Scholar 

  34. Sutphin, G. L. & Kaeberlein, M. Measuring Caenorhabditis elegans life span on solid media. J. Vis. Exp. 27, 1152 (2009)

    Google Scholar 

  35. Gaudet, J. & Mango, S. E. Regulation of organogenesis by the Caenorhabditis elegans FoxA protein PHA-4. Science 295, 821–825 (2002)

    CAS  PubMed  ADS  Google Scholar 

  36. Abada, E. A. et al. C. elegans behavior of preference choice on bacterial food. Mol. Cells 28, 209–213 (2009)

    CAS  PubMed  Google Scholar 

  37. Lomenick, B., Jung, G., Wohlschlegel, J. A. & Huang, J. Target identification using drug affinity responsive target stability (DARTS). Curr. Protoc. Chem. Biol. 3, 163–180 (2011)

    PubMed  PubMed Central  Google Scholar 

  38. Lomenick, B., Olsen, R. W. & Huang, J. Identification of direct protein targets of small molecules. ACS Chem. Biol. 6, 34–46 (2011)

    CAS  PubMed  Google Scholar 

  39. Stubbs, C. J. et al. Application of a proteolysis/mass spectrometry method for investigating the effects of inhibitors on hydroxylase structure. J. Med. Chem. 52, 2799–2805 (2009)

    CAS  PubMed  Google Scholar 

  40. Rogers, G. W. et al. High throughput microplate respiratory measurements using minimal quantities of isolated mitochondria. PLoS ONE 6, e21746 (2011)

    CAS  PubMed  PubMed Central  ADS  Google Scholar 

  41. Alberts, B. Molecular Biology of the Cell 3rd edn (Garland, 1994)

    Google Scholar 

  42. Wu, M. et al. Multiparameter metabolic analysis reveals a close link between attenuated mitochondrial bioenergetic function and enhanced glycolysis dependency in human tumor cells. Am. J. Physiol. Cell Physiol. 292, C125–C136 (2007)

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Pathare, P. P., Lin, A., Bornfeldt, K. E., Taubert, S. & Van Gilst, M. R. Coordinate regulation of lipid metabolism by novel nuclear receptor partnerships. PLoS Genet. 8, e1002645 (2012)

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Pullen, N. & Thomas, G. The modular phosphorylation and activation of p70s6k. FEBS Lett. 410, 78–82 (1997)

    CAS  PubMed  Google Scholar 

  46. Burnett, P. E., Barrow, R. K., Cohen, N. A., Snyder, S. H. & Sabatini, D. M. RAFT1 phosphorylation of the translational regulators p70 S6 kinase and 4E-BP1. Proc. Natl Acad. Sci. USA 95, 1432–1437 (1998)

    CAS  PubMed  ADS  PubMed Central  Google Scholar 

  47. Gingras, A. C. et al. Hierarchical phosphorylation of the translation inhibitor 4E-BP1. Genes Dev. 15, 2852–2864 (2001)

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Sarbassov, D. D., Guertin, D. A., Ali, S. M. & Sabatini, D. M. Phosphorylation and regulation of Akt/PKB by the rictor-mTOR complex. Science 307, 1098–1101 (2005)

    CAS  PubMed  ADS  Google Scholar 

  49. Kim, J., Kundu, M., Viollet, B. & Guan, K. L. AMPK and mTOR regulate autophagy through direct phosphorylation of Ulk1. Nature Cell Biol. 13, 132–141 (2011)

    CAS  PubMed  ADS  Google Scholar 

  50. Kang, C., You, Y. J. & Avery, L. Dual roles of autophagy in the survival of Caenorhabditis elegans during starvation. Genes Dev. 21, 2161–2171 (2007)

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Hansen, M. et al. A role for autophagy in the extension of lifespan by dietary restriction in C. elegans. PLoS Genet. 4, e24 (2008)

    PubMed  PubMed Central  Google Scholar 

  52. Alberti, A., Michelet, X., Djeddi, A. & Legouis, R. The autophagosomal protein LGG-2 acts synergistically with LGG-1 in dauer formation and longevity in C. elegans. Autophagy 6, 622–633 (2010)

    CAS  PubMed  Google Scholar 

  53. Schneider, C. A., Rasband, W. S. & Eliceiri, K. W. NIH Image to ImageJ: 25 years of image analysis. Nature Methods 9, 671–675 (2012)

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Kabeya, Y. et al. LC3, a mammalian homologue of yeast Apg8p, is localized in autophagosome membranes after processing. EMBO J. 19, 5720–5728 (2000)

    CAS  PubMed  PubMed Central  Google Scholar 

  55. MacKenzie, E. D. et al. Cell-permeating α-ketoglutarate derivatives alleviate pseudohypoxia in succinate dehydrogenase-deficient cells. Mol. Cell. Biol. 27, 3282–3289 (2007)

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Zhao, S. et al. Glioma-derived mutations in IDH1 dominantly inhibit IDH1 catalytic activity and induce HIF-1α. Science 324, 261–265 (2009)

    CAS  PubMed  PubMed Central  ADS  Google Scholar 

  57. Xu, W. et al. Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of α-ketoglutarate-dependent dioxygenases. Cancer Cell 19, 17–30 (2011)

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Jin, G. et al. Disruption of wild-type IDH1 suppresses D-2-hydroxyglutarate production in IDH1-mutated gliomas. Cancer Res. 73, 496–501 (2013)

    CAS  PubMed  Google Scholar 

  59. Jung, M. E. & Deng, G. Synthesis of the 1-monoester of 2-ketoalkanedioic acids, for example, octyl α-ketoglutarate. J. Org. Chem. 77, 11002–11005 (2012)

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank S. Lee, M. Hansen, B. Lemire, A. van der Bliek, S. Clarke, T. K. Blackwell, R. Johnson, J. E. Walker, A. G. W. Leslie, K. N. Houk, B. Martin, J. Lusis, J. Gober, Y. Wang and H. Sun for advice and discussions. J. Avruch for the let-363 RNAi vector; J. Powell-Coffman for strains and advice; and K. Yan for technical assistance. Worm strains were provided by the Caenorhabditis Genetics Center, which is funded by the National Institutes of Health (NIH) Office of Research Infrastructure Programs (P40 OD010440). We thank the NIH for traineeship support of R.M.C. (T32 GM007104), M.Y.P. (T32 GM007185), B.L. (T32 GM008496) and M.N. (T32 CA009120). X.F. is a recipient of the China Scholarship Council Scholarship. G.C.M. was supported by Ford Foundation and National Science Foundation Graduate Research Fellowships.

Author information

Authors and Affiliations

Authors

Contributions

Lifespan assays were performed by R.M.C., M.P. and E.H.; DARTS-mass spectrometry by S.D. and B.L.; DARTS-western blots by M.Y.P., H.H. and R.M.C.; mammalian cell experiments by X.F. and H.H.; mitochondrial respiration study design and analyses by L.V. and K.R.; enzyme kinetics and analyses by R.M.C. and J.H.; confocal microscopy by V.S.M., G.C.M. and A.R.F.; ultra-high-performance liquid chromatography-electrospray ionization-tandem mass spectrometry (UHPLC-ESI/MS/MS) by J.X.W. and S.A.T.; compound syntheses by G.D. and M.E.J.; other analyses by H.H., X.F., M.Y.P., D.B., R.M.C., E.H., G.J., G.M.S., C.K. and A.Q. S.A.W., F.F., M.N., A.S.K., H.A.G., H.R. Chang, K.F.F., F.G., M.J., S.A.T., A.S., D.B., H.R. Christofk, C.F.C., M.A.T., M.E.J., L.V., K.R., A.R.F. and M.P. provided guidance, specialized reagents and expertise. J.H. conceived the study. R.M.C. and J.H. wrote the paper. R.M.C., X.F. and J.H. analysed data. All authors discussed the results, commented on the studies and contributed to aspects of preparing the manuscript.

Corresponding author

Correspondence to Jing Huang.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Extended data figures and tables

Extended Data Figure 1 Supplementation with α-KG extends C. elegans adult lifespan but does not change the growth rate of bacteria, or food intake, pharyngeal pumping rate or brood size of the worms.

a, Robust lifespan extension in adult C. elegans by α-KG. 8 mM α-KG increased the mean lifespan of N2 by an average of 47.3% in three independent experiments (P < 0.0001 for every experiment, by log-rank test). Experiment 1, mean lifespan (days of adulthood) with vehicle treatment (mveh) = 18.9 (n = 87 animals tested), mα-KG = 25.8 (n = 96); experiment 2, mveh = 17.5 (n = 119), mα-KG = 25.4 (n = 97); experiment 3, mveh = 16.3 (n = 100), mα-KG = 26.1 (n = 104). b, Worms supplemented with 8 mM α-KG and worms with RNAi knockdown of α-KGDH (encoded by ogdh-1) have increased α-KG levels. Young adult worms were placed on treatment plates seeded with control HT115 E. coli or HT115-expressing ogdh-1 dsRNA, and α-KG content was assayed after 24 h (see Methods). c, α-KG treatment beginning at the egg stage and that beginning in adulthood produced identical lifespan increases. Light red, treatment with vehicle control throughout larval and adult stages (m = 15.6, n = 95); dark red, treatment with vehicle during larval stages and with 8 mM α-KG at adulthood (m = 26.3, n = 102), P < 0.0001 (log-rank test); orange, treatment with 8 mM α-KG throughout larval and adult stages (m = 26.3, n = 102), P < 0.0001 (log-rank test). d, α-KG does not alter the growth rate of the OP50 E. coli, which is the standard laboratory food source for nematodes. α-KG (8 mM) or vehicle (H2O) was added to standard LB media and the pH was adjusted to 6.6 by the addition of NaOH. Bacterial cells from the same overnight OP50 culture were added to the LB ± α-KG mixture at a 1:40 dilution, and then placed in the 37 °C incubator shaker at 300 r.p.m. The absorbance at 595 nm was read at 1 h time intervals to generate the growth curve. e, Schematic representation of food preference assay. f, N2 worms show no preference between OP50 E. coli food treated with vehicle or α-KG (P = 0.85, by t-test, two-tailed, two-sample unequal variance), nor preference between identically treated OP50 E. coli. g, Pharyngeal pumping rate of C. elegans on 8 mM α-KG is not significantly altered (by t-test, two-tailed, two-sample unequal variance). h, Brood size of C. elegans treated with 8 mM α-KG. Brood size analysis was conducted at 20 °C. Ten L4 wild-type worms were each singly placed onto an NGM plate containing vehicle or 8 mM α-KG. Worms were transferred one per plate onto a new plate every day, and the eggs laid were allowed to hatch and develop on the previous plate. Hatchlings were counted as a vacuum was used to remove them from the plate. Animals on 8 mM α-KG showed no significant difference in brood size compared with animals on vehicle plates (P = 0.223, by t-test, two-tailed, two-sample unequal variance). Mean ± s.d. is plotted in all cases.

Source data

Extended Data Figure 2 α-KG binds to the β subunit of ATP synthase and inhibits the activity of complex V but not the other ETC complexes.

a, Western blot showing protection of the ATP-2 protein from Pronase digestion upon α-KG binding in the DARTS assay. The antibody for human ATP5B (Sigma, AV48185) recognizes the epitope 144IMNVIGEPIDERGPIKTKQFAPIHAEAPEFMEMSVEQEILVTGIKVVDLL193 that has 90% identity to the C. elegans ATP-2. The lower molecular weight band near 20 kDa is a proteolytic fragment of the full-length protein corresponding to the domain directly bound by α-KG. b, α-KG does not affect complex IV activity. Complex IV activity was assayed using the MitoTox OXPHOS Complex IV Activity Kit (Abcam, ab109906). Relative complex IV activity was compared to vehicle (H2O) controls. Potassium cyanide (Sigma, 60178) was used as a positive control for the assay. Complex V activity was assayed using the MitoTox Complex V OXPHOS Activity Microplate Assay (Abcam, ab109907). c, atp-2 RNAi worms have lower oxygen consumption compared to control (gfp in RNAi vector), P < 0.0001 (t-test, two-tailed, two-sample unequal variance) for the entire time series (two independent experiments); similar to α-KG-treated worms shown in Fig. 2g. d, α-KG does not affect the electron flow through the ETC. Oxygen consumption rate (OCR) from isolated mouse liver mitochondria at basal (pyruvate and malate as complex I substrate and complex II inhibitor, respectively, in the presence of FCCP) and in response to sequential injection of rotenone (Rote; complex I inhibitor), succinate (Succ; complex II substrate), antimycin A (AA; complex III inhibitor), ascorbate/tetramethylphenylenediamine (Asc/TMPD; cytochrome c (complex IV) substrate). No difference in complex I (C I), complex II (C II) or complex IV (C IV) respiration was observed after 30 min treatment with 800 µM octyl α-KG, whereas complex V was inhibited (see Fig. 2h) by the same treatment (two independent experiments). e, f, No significant difference in coupling (e) or electron flow (f) was observed with either octanol or DMSO vehicle control. g, h, Treatment with 1-octyl α-KG or 5-octyl α-KG gave identical results in coupling (g) or electron flow (h) assays. Mean ± s.d. is plotted in all cases.

Source data

Extended Data Figure 3 Treatment with oligomycin extends C. elegans lifespan and enhances autophagy in a manner dependent on let-363.

a, Oligomycin extends the lifespan of adult C. elegans in a concentration-dependent manner. Treatment with oligomycin began at the young adult stage. 40 µM oligomycin increased the mean lifespan of N2 worms by 32.3% (P < 0.0001, by log-rank test); see Extended Data Table 2 for details. b, Confocal images of GFP::LGG-1 puncta in L3 epidermis of C. elegans with vehicle, oligomycin (40 µM) or α-KG (8 mM), and number of GFP::LGG-1-containing puncta quantified using ImageJ. Bars indicate the mean. Autophagy in C. elegans treated with oligomycin or α-KG is significantly higher than in vehicle-treated control animals (t-test, two-tailed, two-sample unequal variance). c, There is no significant difference (NS) between control worms treated with oligomycin and let-363 RNAi worms treated with vehicle, nor between vehicle- and α-KG-treated let-363 RNAi worms, consistent with independent experiments in Fig. 4b, c; also, oligomycin does not augment autophagy in let-363 RNAi worms (if anything, there may be a small decrease, as indicated by an asterisk); by t-test, two-tailed, two-sample unequal variance. Bars indicate the mean. Photographs were taken at ×100 magnification.

Source data

Extended Data Figure 4 Analyses of oxidative stress in worms treated with α-KG or atp-2 RNAi.

a, The atp-2 RNAi worms have higher levels of 2′,7′-dichlorofluorescein (DCF) fluorescence than gfp control worms (P < 0.0001, by t-test, two-tailed, two-sample unequal variance). Supplementation with α-KG also leads to higher DCF fluorescence, in both HT115- (for RNAi) and OP50-fed worms (P = 0.0007 and P = 0.0012, respectively). Reactive oxygen species (ROS) levels were measured using 2′,7′-dichlorodihydrofluorescein diacetate (H2DCF-DA). As whole worm lysates were used, total cellular oxidative stress was measured here. H2DCF-DA (Molecular Probes, D399) was dissolved in ethanol to a stock concentration of 1.5 mg ml−1. Fresh stock was prepared every time before use. For measuring ROS in worm lysates, a working concentration of H2DCF-DA at 30 ng ml−1 was hydrolysed by 0.1 M NaOH at room temperature for 30 min to generate 2′,7′-dichlorodihydrofluorescein (DCFH) before mixing with whole worm lysates in a black 96-well plate (Greiner Bio-One). Oxidation of DCFH by ROS yields the highly fluorescent DCF. DCF fluorescence was read at excitation/emission of 485/530 nm using SpectraMax MS (Molecular Devices). H2O2 was used as positive control (data not shown). To prepare the worm lysates, synchronized young adult animals were cultivated on plates containing vehicle or 8 mM α-KG and OP50 or HT115 E. coli for 1 day, and then collected and lysed as described in Methods. Mean ± s.d. is plotted. b, There was no significant change in protein oxidation upon α-KG treatment or atp-2 RNAi. Oxidized protein levels were determined by OxyBlot. Synchronized young adult N2 animals were placed onto plates containing vehicle or 8 mM α-KG, and seeded with OP50 or HT115 bacteria that expressed control or atp-2 dsRNA. Adult day 2 and day 3 worms were collected and washed four times with M9 buffer, and then stored at −80 °C for at least 24 h. Laemmli buffer (Biorad, 161-0737) was added to every sample and animals were lysed by alternate boil/freeze cycles. Lysed animals were centrifuged at 14,000 r.p.m. for 10 min at 4 °C to pellet worm debris, and supernatant was collected for OxyBlot analysis. Protein concentration of samples was determined by the 660 nm Protein Assay (Thermo Scientific, 1861426) and normalized for all samples. Carbonylation of proteins in each sample was detected using the OxyBlot Protein Oxidation Detection Kit (Millipore, S7150).

Source data

Extended Data Figure 5 Lifespan extension by α-KG in the absence of aak-2, daf-16, hif-1, vhl-1 or egl-9.

a, Lifespans of α-KG-supplemented N2 worms, mveh = 17.5 (n = 119), mα-KG = 25.4 (n = 97), P < 0.0001; or aak-2(ok524) mutants, mveh = 13.7 (n = 85), mα-KG = 17.1 (n = 83), P < 0.0001. b, N2 worms fed gfp RNAi control, mveh = 18.5 (n = 101), mα-KG = 23.1 (n = 98), P < 0.0001; or daf-16 RNAi, mveh = 14.3 (n = 99), mα-KG = 17.6 (n = 99), P < 0.0001. c, N2 worms, mveh = 21.5 (n = 101), mα-KG = 24.6 (n = 102), P < 0.0001; hif-1(ia7) mutants, mveh = 19.6 (n = 102), mα-KG = 23.6 (n = 101), P < 0.0001; vhl-1(ok161) mutants, mveh = 20.0 (n = 98), mα-KG = 24.9 (n = 100), P < 0.0001; or egl-9(sa307) mutants, mveh = 16.2 (n = 97), mα-KG = 25.6 (n = 96), P < 0.0001. P values were determined by the log-rank test. Number of independent experiments: N2 (8), hif-1 (5), vhl-1 (1) and egl-9 (2); see Extended Data Table 2 for details. Two different hif-1 mutant alleles27 have been used: ia4 (shown in Fig. 3g) is a deletion over several introns and exons; ia7 (shown here) is an early stop codon, causing a truncated protein. Both alleles have the same effect on lifespan27. We tested both alleles for α-KG longevity and obtained the same results.

Source data

Extended Data Figure 6 α-KG decreases TOR pathway activity but does not directly interact with TOR.

a, Phosphorylation of S6K (T389) was decreased in U87 cells treated with octyl α-KG, but not in cells treated with octanol control. The same results were obtained using HEK-293 and MEF cells. b, Phosphorylation of AMPK(T172) is upregulated in WI-38 cells upon complex V inhibition by α-KG, consistent with decreased ATP content in α-KG-treated cells and animals. However, this activation of AMPK appears to require more severe complex V inhibition than the inactivation of mammalian TOR, as either oligomycin or a higher concentration of octyl α-KG was required for increasing phospho (P)-AMPK whereas concentrations of octyl α-KG comparable to those that decreased cellular ATP content (Fig. 2d) or oxygen consumption (Fig. 2f) were also sufficient for decreasing P-S6K. The same results were obtained using U87 cells. Samples were subjected to SDS–PAGE on 4–12% Bis-Tris gradient gel (Invitrogen, NP0322BOX) and western blotted with specific antibodies against P-AMPK T172 (Cell Signaling, 2535S) and AMPK (Cell Signaling, 2603S). c, α-KG still induces autophagy in aak-2 RNAi worms; **P < 0.01 (t-test, two-tailed, two-sample unequal variance). The number of GFP::LGG-1 containing puncta was quantified using ImageJ. Bars indicate the mean. d, e, α-KG does not bind to TOR directly as determined by DARTS. HEK-293 (d) or HeLa (e) cells were lysed in M-PER buffer (Thermo Scientific, 78501) with the addition of protease inhibitors (Roche, 11836153001) and phosphatase inhibitors (50 mM NaF, 10 mM β-glycerophosphate, 5 mM sodium pyrophosphate, 2 mM Na3VO4). Protein concentration of the lysate was measured by BCA Protein Assay kit (Pierce, 23227). Chilled TNC buffer (50 mM Tris-HCl pH 8.0, 50 mM NaCl, 10 mM CaCl2) was added to the protein lysate, and the protein lysate was then incubated with vehicle control (DMSO) or varying concentrations of α-KG for 1 h (d) or 3 h (e) at room temperature. Pronase (Roche, 10165921001) digestions were performed for 20 min at room temperature, and stopped by adding SDS loading buffer and immediately heating at 95 °C for 5 min (d) or 70 °C for 10 min (e). Samples were subjected to SDS–PAGE on 4–12% Bis-Tris gradient gel (Invitrogen, NP0322BOX) and western blotted with specific antibodies against ATP5B (Santa Cruz, sc58618), mammalian TOR (Cell Signaling, 2972) or GAPDH (Ambion, AM4300). ImageJ was used to quantify the mammalian TOR/GAPDH and ATP5B/GAPDH ratios. Susceptibility of the mammalian TOR protein to Pronase digestion is unchanged in the presence of α-KG, whereas, as expected, Pronase resistance in the presence of α-KG is increased for ATP5B, which we identified as a new binding target of α-KG. f, Increased autophagy in HEK-293 cells treated with octyl α-KG was confirmed by western blot analysis of MAP1 LC3 (Novus, NB100-2220), consistent with decreased phosphorylation of the autophagy-initiating kinase ULK1 (Fig. 4a).

Source data

Extended Data Figure 7 Autophagy is enhanced in C. elegans treated with ogdh-1 RNAi.

a, Confocal images of GFP::LGG-1 puncta in the epidermis of mid-L3 stage, control or ogdh-1 knockdown C. elegans treated with vehicle or α-KG (8 mM). b, Number of GFP::LGG-1 puncta quantified using ImageJ. Bars indicate the mean. ogdh-1 RNAi worms have significantly higher autophagy levels, and α-KG does not significantly augment autophagy in ogdh-1 RNAi worms (t-test, two-tailed, two-sample unequal variance). Photographs were taken at ×100 magnification.

Source data

Extended Data Table 1 Enriched proteins in the α-KG DARTS sample
Extended Data Table 2 Summary of lifespan data

Supplementary information

Supplementary Information

This file contains Supplementary Notes and additional references. (PDF 242 kb)

Example of a vehicle-treated day 16 adult animal

The animal had lost all motility in the body and could only move its head slowly. (WMV 2122 kb)

Example of a α-KG-treated day 16 adult animal

The animal remained youthful and exhibited full body movements. (WMV 2200 kb)

PowerPoint slides

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chin, R., Fu, X., Pai, M. et al. The metabolite α-ketoglutarate extends lifespan by inhibiting ATP synthase and TOR. Nature 510, 397–401 (2014). https://doi.org/10.1038/nature13264

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature13264

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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