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:

Mitochondria and cancer

Key Points

  • Warburg observed, 70 years ago, that tumours produce excess lactate in the presence of oxygen. This became known as aerobic glycolysis or the 'Warburg effect' which he interpreted as mitochondrial dysfunction. However, it is now clear that mitochondrial function is essential for cancer cell viability, because elimination of cancer cell mitochondrial DNAs (mtDNAs) reduces their growth rate and tumorigenicity.

  • The mitochondrial genome encompasses thousands of copies of the mtDNA and more than one thousand nuclear DNA (nDNA)-encoded genes. mtDNA mutations have been found in various cancers and seem to alter mitochondrial metabolism, enhance tumorigenesis and permit cancer cell adaptation to changing environments.

  • Mutations in nDNA genes involved in mitochondrial metabolism, including succinate dehydrogenase (SDH), fumarate hydratase (FH), isocitrate dehydrogenase 1 (IDH1) and IDH2, result in increased succinate, fumarate, or R(-)-2-hydroxyglutarate levels. These metabolic alterations can inhibit various α-ketoglutarate-dependent dioxygenases; they can also activate the NFE2-related factor 2 (NRF2) stress response pathway. All of these effects can contribute to tumorigenesis.

  • Activation of signalling pathways and oncogenes that are known to be important in tumorigenesis also affect mitochondrial function. The PI3K–PTEN–AKT pathway shifts metabolism from oxidative to glycolytic, thus permitting the redistribution of glycolytic nutrients from catabolism to anabolism. Activation of MYC induces glutaminolysis, which provides anaplerotic substrates to the mitochondrial tricarboxylic acid cycle, thus enhancing citrate production and its export to the cytosol to provide acetyl-CoA for lipid biogenesis and protein modifications.

  • Altered mitochondrial metabolism can increase the production of mitochondrial reactive oxygen species (ROS) and change the cellular redox status, thus altering the activities of transcription factors such as HIF1α and FOS–JUN to change gene expression and stimulate cancer cell proliferation.

  • A decrease of the mitochondrial membrane potential or mutation of the promyelocytic leukaemia (PML) gene reduces mitochondrial Ca2+ uptake, thus decreasing the activation of the mitochondrial intrinsic apoptosis pathway.

  • Reduced mitochondrial Ca2+ retention increases the cytosolic Ca2+ concentration. This activates mitochondrial retrograde signalling through stimulation of calcineurin and IκBβ-dependent NF-κB, activation of enhanceosome-driven transcription and increased metastatic potential.

  • Cancer cell ROS production inactivates caveolin 1 in adjacent stromal fibroblasts. This increases mitophagy, reduces mitochondrial function and increases lactate production in these fibroblasts. Secreted stromal cell lactate then fuels cancer cell oxidative metabolism, which drives tumour growth and proliferation. This is known as the 'reverse Warburg effect'.

Abstract

Contrary to conventional wisdom, functional mitochondria are essential for the cancer cell. Although mutations in mitochondrial genes are common in cancer cells, they do not inactivate mitochondrial energy metabolism but rather alter the mitochondrial bioenergetic and biosynthetic state. These states communicate with the nucleus through mitochondrial 'retrograde signalling' to modulate signal transduction pathways, transcriptional circuits and chromatin structure to meet the perceived mitochondrial and nuclear requirements of the cancer cell. Cancer cells then reprogramme adjacent stromal cells to optimize the cancer cell environment. These alterations activate out-of-context programmes that are important in development, stress response, wound healing and nutritional status.

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: Mitochondrial genome and mitochondrial biogenesis.
Figure 2: Mitochondrial bioenergetics and cancer cell mutations.
Figure 3: Mitochondrial physiology.
Figure 4: The mitochondrial NADPH shuttle system, and IDH 1 and IDH2 mutations

Similar content being viewed by others

References

  1. Warburg, O. The Metabolism of Tumors (R. R. Smith, 1931).

    Google Scholar 

  2. Warburg, O. On the origin of cancer cells. Science 123, 309–314 (1956).

    Article  CAS  PubMed  Google Scholar 

  3. Wallace, D. C. Mitochondria and cancer: Warburg addressed. Cold Spring Harb. Symp. Quant. Biol. 70, 363–374 (2005).

    Article  CAS  PubMed  Google Scholar 

  4. Lin, M. Molecular imaging using positron emission tomography in colorectal cancer. Discov. Med. 11, 435–447 (2011).

    PubMed  Google Scholar 

  5. Pedersen, P. L. Tumor mitochondria and the bioenergetics of cancer cells. Prog . Exp. Tumor Res. 22, 190–274 (1978).

    Article  CAS  PubMed  Google Scholar 

  6. Lane, N. & Martin, W. The energetics of genome complexity. Nature 467, 929–934 (2010).

    Article  CAS  PubMed  Google Scholar 

  7. Wallace, D. C. Why do we have a maternally inherited mitochondrial DNA?Insights from evolutionary medicine. Annu. Rev. Biochem. 76, 781–821 (2007).

    Article  CAS  PubMed  Google Scholar 

  8. Horton, T. M. et al. Novel mitochondrial DNA deletion found in a renal cell carcinoma. Genes Chromosomes Cancer 15, 95–101 (1996).

    Article  CAS  PubMed  Google Scholar 

  9. Desjardins, P., Frost, E. & Morais, R. Ethidium bromide-induced loss of mitochondrial DNA from primary chicken embryo fibroblasts. Mol. Cell. Biol. 5, 1163–1169 (1985).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Desjardins, P., de Muys, J. M. & Morais, R. An established avian fibroblast cell line without mitochondrial DNA. Somat. Cell Genet. 12, 133–139 (1986).

    Article  CAS  Google Scholar 

  11. King, M. P. & Attardi, G. Human cells lacking mtDNA: repopulation with exogenous mitochondria by complementation. Science 246, 500–503 (1989).

    Article  CAS  PubMed  Google Scholar 

  12. Magda, D. et al. mtDNA depletion confers specific gene expression profiles in human cells grown in culture and in xenograft. BMC Genomics 9, 521 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  13. Morais, R. et al. Tumor-forming ability in athymic nude mice of human cell lines devoid of mitochondrial DNA. Cancer Res. 54, 3889–3896 (1994).

    CAS  PubMed  Google Scholar 

  14. Cavalli, L. R., Varella-Garcia, M. & Liang, B. C. Diminished tumorigenic phenotype after depletion of mitochondrial DNA. Cell Growth Differ. 8, 1189–1198 (1997).

    CAS  PubMed  Google Scholar 

  15. Weinberg, F. et al. Mitochondrial metabolism and ROS generation are essential for Kras-mediated tumorigenicity. Proc. Natl Acad. Sci. USA 107, 8788–8793 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Murgia, M., Giorgi, C., Pinton, P. & Rizzuto, R. Controlling metabolism and cell death: at the heart of mitochondrial calcium signalling. J. Mol. Cell. Cardiol. 46, 781–788 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Aanen, D. K. & Maas, M. F. Recruitment of healthy mitochondria fuels transmissible cancers. Trends Genet. 28, 1–6 (2012).

    Article  CAS  PubMed  Google Scholar 

  18. Rebbeck, C. A., Leroi, A. M. & Burt, A. Mitochondrial capture by a transmissible cancer. Science 331, 303 (2011).

    Article  CAS  PubMed  Google Scholar 

  19. Brandon, M., Baldi, P. & Wallace, D. C. Mitochondrial mutations in cancer. Oncogene 25, 4647–4662 (2006).

    Article  CAS  PubMed  Google Scholar 

  20. Chinnery, P. F., Samuels, D. C., Elson, J. & Turnbull, D. M. Accumulation of mitochondrial DNA mutations in ageing, cancer, and mitochondrial disease: is there a common mechanism? Lancet 360, 1323–1325 (2002).

    Article  CAS  PubMed  Google Scholar 

  21. Copeland, W. C., Wachsman, J. T., Johnson, F. M. & Penta, J. S. Mitochondrial DNA alterations in cancer. Cancer Invest. 20, 557–569 (2002).

    Article  CAS  PubMed  Google Scholar 

  22. Gasparre, G. et al. Clonal expansion of mutated mitochondrial DNA is associated with tumor formation and complex I deficiency in the benign renal oncocytoma. Hum. Mol. Genet. 17, 986–995 (2008).

    Article  CAS  PubMed  Google Scholar 

  23. Bartoletti-Stella, A. et al. Mitochondrial DNA mutations in oncocytic adnexal lacrimal glands of the conjunctiva. Arch. Ophthalmol. 129, 664–666 (2011).

    Article  CAS  PubMed  Google Scholar 

  24. Pereira, L., Soares, P., Maximo, V. & Samuels, D. C. Somatic mitochondrial DNA mutations in cancer escape purifying selection and high pathogenicity mutations lead to the oncocytic phenotype: pathogenicity analysis of reported somatic mtDNA mutations in tumors. BMC Cancer 12, 53 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Salas, A. et al. A critical reassessment of the role of mitochondria in tumorigenesis. PLoS Med. 2, e296 (2005).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Meierhofer, D. et al. Mitochondrial DNA mutations in renal cell carcinomas revealed no general impact on energy metabolism. Br. J. Cancer 94, 268–274 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Czarnecka, A. M. et al. Molecular oncology focus - is carcinogenesis a 'mitochondriopathy'? J. Biomed. Sci. 17, 31 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. Howell, A. N. & Sager, R. Tumorigenicity and its suppression in cybrids of mouse and Chinese hamster cell lines. Proc. Natl Acad. Sci. USA 75, 2358–2362 (1978).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Petros, J. A. et al. mtDNA mutations increase tumorigenicity in prostate cancer. Proc. Natl Acad. Sci. USA 102, 719–724 (2005). A demonstration that human mtDNA mutations that increase ROS production enhance tumorigenesis, whereas normal mtDNAs suppress tumorigenesis.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Shidara, Y. et al. Positive contribution of pathogenic mutations in the mitochondrial genome to the promotion of cancer by prevention from apoptosis. Cancer Res. 65, 1655–1663 (2005).

    Article  CAS  PubMed  Google Scholar 

  31. Ishikawa, K. et al. ROS-generating mitochondrial DNA mutations can regulate tumor cell metastasis. Science 320, 661–664 (2008). This study shows that mouse mtDNA mutations that increase mitochondrial ROS levels also increase tumorigenesis.

    Article  CAS  PubMed  Google Scholar 

  32. Canter, J. A., Kallianpur, A. R., Parl, F. F. & Millikan, R. C. Mitochondrial DNA G10398A polymorphism and invasive breast cancer in African-American women. Cancer Res. 65, 8028–8033 (2005).

    Article  CAS  PubMed  Google Scholar 

  33. Liu, V. W. et al. Mitochondrial DNA variant 16189T>C is associated with susceptibility to endometrial cancer. Hum. Mut. 22, 173–174 (2003).

    Article  CAS  PubMed  Google Scholar 

  34. Permuth-Wey, J. et al. Inherited variants in mitochondrial biogenesis genes may influence epithelial ovarian cancer risk. Cancer Epidemiol. Biomarkers Prev. 20, 1131–1145 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Zhang, J. et al. Strikingly higher frequency in centenarians and twins of mtDNA mutation causing remodeling of replication origin in leukocytes. Proc. Natl Acad. Sci. USA 100, 1116–1121 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Zhai, K., Chang, L., Zhang, Q., Liu, B. & Wu, Y. Mitochondrial C150T polymorphism increases the risk of cervical cancer and HPV infection. Mitochondrion 11, 559–563 (2011).

    Article  CAS  PubMed  Google Scholar 

  37. Wallace, D. C. Bioenergetic origins of complexity and diseases. Cold Spring Harb. Symp. Quant. Biol. 76, 1–16 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Ruiz-Pesini, E. & Wallace, D. C. Evidence for adaptive selection acting on the tRNA and rRNA genes of the human mitochondrial DNA. Hum. Mut. 27, 1072–1081 (2006).

    Article  CAS  PubMed  Google Scholar 

  39. Mishmar, D. et al. Natural selection shaped regional mtDNA variation in humans. Proc. Natl Acad. Sci. USA 100, 171–176 (2003).

    Article  CAS  PubMed  Google Scholar 

  40. Wallace, D. C. Colloquium paper: bioenergetics, the origins of complexity, and the ascent of man. Proc. Natl Acad. Sci. USA 107, 8947–8953 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Parrella, P. et al. Detection of mitochondrial DNA mutations in primary breast cancer and fine-needle aspirates. Cancer Res. 61, 7623–7626 (2001).

    CAS  PubMed  Google Scholar 

  42. Guo, J. et al. Frequent truncating mutation of TFAM induces mitochondrial DNA depletion and apoptotic resistance in microsatellite-unstable colorectal cancer. Cancer Res. 71, 2978–2987 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Han, B. et al. Human mitochondrial transcription factor A functions in both nuclei and mitochondria and regulates cancer cell growth. Biochem. Biophys. Res. Commun. 408, 45–51 (2011).

    Article  CAS  PubMed  Google Scholar 

  44. Bogenhagen, D. F., Rousseau, D. & Burke, S. The layered structure of human mitochondrial DNA nucleoids. J. Biol. Chem. 283, 3665–3675 (2008).

    Article  CAS  PubMed  Google Scholar 

  45. Bogenhagen, D. F. Mitochondrial DNA nucleoid structure. Biochim. Biophys. Acta 1819, 914–920 (2011).

    Article  PubMed  CAS  Google Scholar 

  46. Khidr, L. et al. Role of SUV3 helicase in maintaining mitochondrial homeostasis in human cells. J. Biol. Chem. 283, 27064–27073 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Chen, P.-L. et al. Mitochondrial genome instability resulting from SUV3 haploinsufficiency leads to tumorigenesis and shortened lifespan. Oncogene 7 May 2012 (doi:10.1038/onc.2012.120).

  48. Bardella, C., Pollard, P. J. & Tomlinson, I. SDH mutations in cancer. Biochim. Biophys. Acta 1807, 1432–1443 (2011).

    Article  CAS  PubMed  Google Scholar 

  49. Baysal, B. E. et al. Mutations in SDHD, a mitochondrial complex II gene, in hereditary paraganglioma. Science 287, 848–851 (2000). The first report that inactivation of SDHD can cause paragangliosis.

    Article  CAS  PubMed  Google Scholar 

  50. Niemann, S. & Muller, U. Mutations in SDHC cause autosomal dominant paraganglioma, type 3. Nature Genet. 26, 268–270 (2000).

    Article  CAS  PubMed  Google Scholar 

  51. Astuti, D. et al. Gene mutations in the succinate dehydrogenase subunit SDHB cause susceptibility to familial pheochromocytoma and to familial paraganglioma. Am. J. Hum. Genet. 69, 49–54 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Kurelac, I., Romeo, G. & Gasparre, G. Mitochondrial metabolism and cancer. Mitochondrion 11, 635–637 (2011).

    Article  PubMed  CAS  Google Scholar 

  53. Wallace, D. C. & Fan, W. Energetics, epigenetics, mitochondrial genetics. Mitochondrion 10, 12–31 (2010).

    Article  CAS  PubMed  Google Scholar 

  54. Chandel, N. S. et al. Reactive oxygen species generated at mitochondrial complex III stabilize hypoxia-inducible factor-1α during hypoxia: a mechanism of O2 sensing. J. Biol. Chem. 275, 25130–25138 (2000). Evidence that increased mitochondrial ROS levels can inactivate PHDs and activate HIF1α.

    Article  CAS  PubMed  Google Scholar 

  55. Guzy, R. D., Sharma, B., Bell, E., Chandel, N. S. & Schumacker, P. T. Loss of the SdhB, but Not the SdhA, subunit of complex II triggers reactive oxygen species-dependent hypoxia-inducible factor activation and tumorigenesis. Mol. Cell. Biochem. 28, 718–731 (2008).

    Article  CAS  Google Scholar 

  56. Selak, M. A., Duran, R. V. & Gottlieb, E. Redox stress is not essential for the pseudo-hypoxic phenotype of succinate dehydrogenase deficient cells. Biochim. Biophys. Acta 1757, 567–572 (2006).

    Article  CAS  PubMed  Google Scholar 

  57. Xiao, M. et al. Inhibition of α-KG-dependent histone and DNA demethylases by fumarate and succinate that are accumulated in mutations of FH and SDH tumor suppressors. Genes Dev. 26, 1326–1338 (2012). A demonstration that excessive dicarboxylic acids from SDH and FH inactivation inhibit α-ketoglutarate-dependent dioxygenases, thus altering chromatin structure and gene expression.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Picaud, S. et al. Structural basis of fumarate hydratase deficiency. J. Inherit. Metab. Dis. 34, 671–676 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Frezza, C. et al. Haem oxygenase is synthetically lethal with the tumour suppressor fumarate hydratase. Nature 477, 225–228 (2011). A report that increased fumarate levels activates the NRF2 stress response, inducing HMOX1 and haem degradation.

    Article  CAS  PubMed  Google Scholar 

  60. Adam, J. et al. Renal cyst formation in Fh1-deficient mice is independent of the Hif/Phd pathway: roles for fumarate in KEAP1 succination and Nrf2 signaling. Cancer Cell 20, 524–537 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Stark, R. et al. Phosphoenolpyruvate cycling via mitochondrial phosphoenolpyruvate carboxykinase links anaplerosis and mitochondrial GTP with insulin secretion. J. Biol. Chem. 284, 26578–26590 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Kibbey, R. G. et al. Mitochondrial GTP regulates glucose-stimulated insulin secretion. Cell Metab. 5, 253–264 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Thompson, C. B. Metabolic enzymes as oncogenes or tumor suppressors. New Engl. J. Med. 360, 813–815 (2009).

    Article  CAS  PubMed  Google Scholar 

  64. Ward, P. S. et al. The common feature of leukemia-associated IDH1 and IDH2 mutations is a neomorphic enzyme activity converting α-ketoglutarate to 2-hydroxyglutarate. Cancer Cell 17, 225–234 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Ward, P. S. et al. Identification of additional IDH mutations associated with oncometabolite R(–)-2-hydroxyglutarate production. Oncogene 31, 2491–2498 (2012).

    Article  CAS  PubMed  Google Scholar 

  66. Dang, L. et al. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature 462, 739–744 (2009). This study reports that heterozygous IDH1 mutation creates a neomorphic enzyme that generates the novel metabolite ( R )-2HG.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Koivunen, P. et al. Transformation by the (R)-enantiomer of 2-hydroxyglutarate linked to EGLN activation. Nature 483, 484–488 (2012). This study shows that in contrast to succinate and fumarate, ( R )-2HG does not inactivate PHDs and activate HIF1α, implying that tumorigenesis involves an alternative pathway.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Figueroa, M. E. et al. DNA methylation signatures identify biologically distinct subtypes in acute myeloid leukemia. Cancer Cell 17, 13–27 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Chowdhury, R. et al. The oncometabolite 2-hydroxyglutarate inhibits histone lysine demethylases. EMBO Rep. 12, 463–469 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Lu, C. et al. IDH mutation impairs histone demethylation and results in a block to cell differentiation. Nature 483, 474–478 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Turcan, S. et al. IDH1 mutation is sufficient to establish the glioma hypermethylator phenotype. Nature 483, 479–483 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Yoon, J. C. et al. Wnt signaling regulates mitochondrial physiology and insulin sensitivity. Genes Dev. 24, 1507–1518 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Bjornsson, H. T. et al. Epigenetic specificity of loss of imprinting of the IGF2 gene in Wilms tumors. J. Natl Cancer Inst. 99, 1270–1273 (2007).

    Article  CAS  PubMed  Google Scholar 

  74. Feinberg, A. P. Phenotypic plasticity and the epigenetics of human disease. Nature 447, 433–440 (2007).

    Article  CAS  PubMed  Google Scholar 

  75. Feinberg, A. P. Epigenetics at the epicenter of modern medicine. JAMA 299, 1345–1350 (2008).

    Article  CAS  PubMed  Google Scholar 

  76. Smiraglia, D. J., Kulawiec, M., Bistulfi, G. L., Gupta, S. G. & Singh, K. K. A novel role for mitochondria in regulating epigenetic modification in the nucleus. Cancer Biol. Ther. 7, 1182–1190 (2008).

    Article  CAS  PubMed  Google Scholar 

  77. Naviaux, R. K. Mitochondrial control of epigenetics. Cancer Biol. Ther. 7, 1191–1193 (2008).

    Article  CAS  PubMed  Google Scholar 

  78. Toye, A. A. et al. A genetic and physiological study of impaired glucose homeostasis control in C57BL/6J mice. Diabetologia 48, 675–686 (2005).

    Article  CAS  PubMed  Google Scholar 

  79. Freeman, H. C., Hugill, A., Dear, N. T., Ashcroft, F. M. & Cox, R. D. Deletion of nicotinamide nucleotide transhydrogenase: a new quantitive trait locus accounting for glucose intolerance in C57BL/6J mice. Diabetes 55, 2153–2156 (2006).

    Article  CAS  PubMed  Google Scholar 

  80. Collins, S., Martin, T. L., Surwit, R. S. & Robidoux, J. Genetic vulnerability to diet-induced obesity in the C57BL/6J mouse: physiological and molecular characteristics. Physiol. Behav. 81, 243–248 (2004).

    Article  CAS  PubMed  Google Scholar 

  81. Nicholson, A. et al. Diet-induced obesity in two C57BL/6 substrains with intact or mutant nicotinamide nucleotide transhydrogenase (Nnt) gene. Obesity 18, 1902–1905 (2010).

    Article  CAS  PubMed  Google Scholar 

  82. Huang, T. T. et al. Genetic modifiers of the phenotype of mice deficient in mitochondrial superoxide dismutase. Hum. Mol. Genet. 15, 1187–1194 (2006).

    Article  CAS  PubMed  Google Scholar 

  83. Mullen, A. R. et al. Reductive carboxylation supports growth in tumour cells with defective mitochondria. Nature 481, 385–388 (2012). This study shows that mitochondrial α-ketoglutarate can be reductively carboxylated using mitochondrial NADPH to increase citrate production.

    Article  CAS  Google Scholar 

  84. 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 

  85. Melov, S. et al. Mitochondrial disease in superoxide dismutase 2 mutant mice. Proc. Natl Acad. Sci. USA 96, 846–851 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Yan, L. J., Levine, R. L. & Sohal, R. S. Oxidative damage during aging targets mitochondrial aconitase. Proc. Natl Acad. Sci. USA 94, 11168–11172 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Tong, J., Schriner, S. E., McCleary, D., Day, B. J. & Wallace, D. C. Life extension through neurofibromin mitochondrial regulation and antioxidant therapy for Neurofibromatosis-1 in Drosophila melanogaster. Nature Genet. 39, 476–485 (2007).

    Article  CAS  PubMed  Google Scholar 

  88. Sasaki, M. et al. IDH1(R132H) mutation increases murine haematopoietic progenitors and alters epigenetics. Nature 488, 656–659 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Gupta, S. C. et al. Upsides and downsides of ROS for cancer: the roles of ROS in tumorigenesis, prevention, and therapy. Antioxid. Redox Signal. 16, 1295–1322 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Burdon, R. H. Superoxide and hydrogen peroxide in relation to mammalian cell proliferation. Free Rad. Biol. Med. 18, 775–794 (1995).

    Article  CAS  PubMed  Google Scholar 

  91. Lander, H. M. An essential role for free radicals and derived species in signal transduction. FASEB J. 11, 118–124 (1997).

    Article  CAS  PubMed  Google Scholar 

  92. Abate, C., Patel, L., Rauscher, F. J. & Curran, T. Redox regulation of fos and jun DNA-binding activity in vitro. Science 249, 1157–1161 (1990). The first demonstration that FOS and JUN are regulated by cysteine oxidation–reduction.

    Article  CAS  PubMed  Google Scholar 

  93. Liu, H., Colavitti, R., Rovira, I. I. & Finkel, T. Redox-dependent transcriptional regulation. Circul. Res. 97, 967–974 (2005).

    Article  CAS  Google Scholar 

  94. Ordway, J. M., Eberhart, D. & Curran, T. Cysteine 64 of Ref-1 is not essential for redox regulation of AP-1 DNA binding. Mol. Cell. Biol. 23, 4257–4266 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Xanthoudakis, S., Miao, G. G. & Curran, T. The redox and DNA-repair activities of Ref-1 are encoded by nonoverlapping domains. Proc. Natl Acad. Sci. USA 91, 23–27 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Go, Y. M. & Jones, D. P. Redox compartmentalization in eukaryotic cells. Biochim. Biophys. Acta 1780, 1273–1290 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Jones, D. P. Radical-free biology of oxidative stress. Am. J. Physiol. Cell Physiol. 295, C849–C868 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Kemp, M., Go, Y. M. & Jones, D. P. Nonequilibrium thermodynamics of thiol/disulfide redox systems: a perspective on redox systems biology. Free Rad. Biol. Med. 44, 921–937 (2008).

    Article  CAS  PubMed  Google Scholar 

  99. Guzy, R. D., Mack, M. M. & Schumacker, P. T. Mitochondrial complex III is required for hypoxia-induced ROS production and gene transcription in yeast. Antioxid. Redox Signal. 9, 1317–1328 (2007).

    Article  CAS  PubMed  Google Scholar 

  100. Sanchez-Cenizo, L. et al. Up-regulation of the ATPase inhibitory factor 1 (IF1) of the mitochondrial H+-ATP synthase in human tumors mediates the metabolic shift of cancer cells to a Warburg phenotype. J. Biol. Chem. 285, 25308–25313 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Willers, I. M. & Cuezva, J. M. Post-transcriptional regulation of the mitochondrial H+-ATP synthase: a key regulator of the metabolic phenotype in cancer. Biochim. Biophys. Acta 1807, 543–551 (2011).

    Article  CAS  PubMed  Google Scholar 

  102. Jones, R. G. & Thompson, C. B. Tumor suppressors and cell metabolism: a recipe for cancer growth. Genes Dev. 23, 537–548 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Wise, D. R. & Thompson, C. B. Glutamine addiction: a new therapeutic target in cancer. Trends Biochem. Sci. 35, 427–433 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. DeBerardinis, R. J., Lum, J. J., Hatzivassiliou, G. & Thompson, C. B. The biology of cancer: metabolic reprogramming fuels cell growth and proliferation. Cell Metab. 7, 11–20 (2008). A report that activation of the PI3K–PTEN–AKT pathway redirects cellular metabolism from oxidative catabolism to glycolytic anabolism, thus enhancing cancer cell biogenesis.

    Article  CAS  PubMed  Google Scholar 

  105. Pedersen, P. L., Mathupala, S., Rempel, A., Geschwind, J. F. & Ko, Y. H. Mitochondrial bound type II hexokinase: a key player in the growth and survival of many cancers and an ideal prospect for therapeutic intervention. Biochim. Biophys. Acta 1555, 14–20 (2002).

    Article  CAS  PubMed  Google Scholar 

  106. Gatenby, R. A. & Gillies, R. J. Why do cancers have high aerobic glycolysis? Nature Rev. Cancer 4, 891–899 (2004).

    Article  CAS  Google Scholar 

  107. Nemoto, S. & Finkel, T. Ageing and the mystery at Arles. Nature 429, 149–152 (2004).

    Article  CAS  PubMed  Google Scholar 

  108. Wallace, D. C. A mitochondrial paradigm of metabolic and degenerative diseases, aging, and cancer: a dawn for evolutionary medicine. Annu. Rev. Genet. 39, 359–407 (2005). An overview of mitochondrial biology and genetics and their relation to disease.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Daitoku, H., Yamagata, K., Matsuzaki, H., Hatta, M. & Fukamizu, A. Regulation of PGC-1 promoter activity by protein kinase B and the forkhead transcription factor FKHR. Diabetes 52, 642–649 (2003).

    Article  CAS  PubMed  Google Scholar 

  110. Spiegelman, B. M. & Heinrich, R. Biological control through regulated transcriptional coactivators. Cell 119, 157–167 (2004).

    Article  CAS  PubMed  Google Scholar 

  111. Ferber, E. C. et al. FOXO3a regulates reactive oxygen metabolism by inhibiting mitochondrial gene expression. Cell Death Differ. 19, 968–979 (2012).

    Article  CAS  PubMed  Google Scholar 

  112. Li, F. et al. Myc stimulates nuclearly encoded mitochondrial genes and mitochondrial biogenesis. Mol. Cell. Biol. 25, 6225–6234 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Bluemlein, K. et al. No evidence for a shift in pyruvate kinase PKM1 to PKM2 expression during tumorigenesis. Oncotarget 2, 393–400 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  114. Hitosugi, T. et al. Tyrosine phosphorylation inhibits PKM2 to promote the Warburg effect and tumor growth. Sci. Signal. 2, ra73 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  115. Anastasiou, D. et al. Inhibition of pyruvate kinase M2 by reactive oxygen species contributes to cellular antioxidant responses. Science 334, 1278–1283 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Gruning, N. M. et al. Pyruvate kinase triggers a metabolic feedback loop that controls redox metabolism in respiring cells. Cell Metab. 14, 415–427 (2011). A demonstration that oxidation of PKM2 inhibits glycolysis and redirects substrates into the pentose phosphate pathway to synthesize NADPH and enhance antioxidant defences.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  117. Hamanaka, R. B. & Chandel, N. S. Warburg effect and redox balance. Science 334, 1219–1220 (2011).

    Article  CAS  PubMed  Google Scholar 

  118. Bricker, D. K. et al. A mitochondrial pyruvate carrier required for pyruvate uptake in yeast, Drosophila, and humans. Science 337, 96–100 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Hoyos, B., Acin-Perez, R., Fischman, D. A., Manfredi, G. & Hammerling, U. Hiding in plain sight: uncovering a new function of vitamin A in redox signaling. Biochim. Biophys. Acta 1821, 241–247 (2012).

    Article  CAS  PubMed  Google Scholar 

  120. Zaugg, K. et al. Carnitine palmitoyltransferase 1C promotes cell survival and tumor growth under conditions of metabolic stress. Genes Dev. 25, 1041–1051 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. DeBerardinis, R. J. et al. Beyond aerobic glycolysis: transformed cells can engage in glutamine metabolism that exceeds the requirement for protein and nucleotide synthesis. Proc. Natl Acad. Sci. USA 104, 19345–19350 (2007). This study shows that MYC induction of glutaminolysis provides anaplerotic TCA cycle intermediates to generate citrate and sustain cytosolic fatty acid synthesis.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Wise, D. R. et al. Myc regulates a transcriptional program that stimulates mitochondrial glutaminolysis and leads to glutamine addiction. Proc. Natl Acad. Sci. USA 105, 18782–18787 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Bensaad, K., Cheung, E. C. & Vousden, K. H. Modulation of intracellular ROS levels by TIGAR controls autophagy. EMBO J. 28, 3015–3026 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Bensaad, K. et al. TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell 126, 107–120 (2006).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  126. Sahin, E. & Depinho, R. A. Linking functional decline of telomeres, mitochondria and stem cells during ageing. Nature 464, 520–528 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Fukuda, R. et al. HIF-1 regulates cytochrome oxidase subunits to optimize efficiency of respiration in hypoxic cells. Cell 129, 111–122 (2007).

    Article  CAS  PubMed  Google Scholar 

  128. Semenza, G. L. Oxygen-dependent regulation of mitochondrial respiration by hypoxia-inducible factor 1. Biochem. J. 405, 1–9 (2007).

    Article  CAS  PubMed  Google Scholar 

  129. Zhang, H. et al. HIF-1 inhibits mitochondrial biogenesis and cellular respiration in VHL-deficient renal cell carcinoma by repression of C-MYC activity. Cancer Cell 11, 407–420 (2007).

    Article  CAS  PubMed  Google Scholar 

  130. Zhang, H. et al. Mitochondrial autophagy is an HIF-1-dependent adaptive metabolic response to hypoxia. J. Biol. Chem. 283, 10892–10903 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Semenza, G. L. Mitochondrial autophagy: life and breath of the cell. Autophagy 4, 534–536 (2008).

    Article  CAS  PubMed  Google Scholar 

  132. Devlin, C., Greco, S., Martelli, F. & Ivan, M. miR-210: more than a silent player in hypoxia. IUBMB Life 63, 94–100 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Bell, E. L., Emerling, B. M., Ricoult, S. J. & Guarente, L. SirT3 suppresses hypoxia inducible factor 1α and tumor growth by inhibiting mitochondrial ROS production. Oncogene 30, 2986–2996 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Yang, J. et al. Human CHCHD4 mitochondrial proteins regulate cellular oxygen consumption rate and metabolism and provide a critical role in hypoxia signaling and tumor progression. J. Clin. Invest. 122, 600–611 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Luo, W. et al. Pyruvate kinase M2 is a PHD3-stimulated coactivator for hypoxia-inducible factor 1. Cell 145, 732–744 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Luo, W. & Semenza, G. L. Pyruvate kinase M2 regulates glucose metabolism by functioning as a coactivator for hypoxia-inducible factor 1 in cancer cells. Oncotarget 2, 551–556 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  137. McCormack, J. G. & Denton, R. M. A comparative study of the regulation of Ca2+ of the activities of the 2-oxoglutarate dehydrogenase complex and NAD+-isocitrate dehydrogenase from a variety of sources. Biochem. J. 196, 619–624 (1981).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. McCormack, J. G., Halestrap, A. P. & Denton, R. M. Role of calcium ions in regulation of mammalian intramitochondrial metabolism. Physiol. Rev. 70, 391–425 (1990).

    Article  CAS  PubMed  Google Scholar 

  139. Baughman, J. M. et al. Integrative genomics identifies MCU as an essential component of the mitochondrial calcium uniporter. Nature 476, 341–345 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. De Stefani, D., Raffaello, A., Teardo, E., Szabo, I. & Rizzuto, R. A forty-kilodalton protein of the inner membrane is the mitochondrial calcium uniporter. Nature 476, 336–340 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Pinton, P., Giorgi, C. & Pandolfi, P. P. The role of PML in the control of apoptotic cell fate: a new key player at ER-mitochondria sites. Cell Death Differ. 18, 1450–1456 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Martinez-Caballero, S. et al. Assembly of the mitochondrial apoptosis-induced channel, MAC. J. Biol. Chem. 284, 12235–12245 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Peixoto, P. M., Ryu, S. Y., Bombrun, A., Antonsson, B. & Kinnally, K. W. MAC inhibitors suppress mitochondrial apoptosis. Biochem. J. 423, 381–387 (2009).

    Article  CAS  PubMed  Google Scholar 

  144. Dejean, L. M. et al. MAC and Bcl-2 family proteins conspire in a deadly plot. Biochim. Biophys. Acta 1797, 1231–1238 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Amuthan, G. et al. Mitochondria-to-nucleus stress signaling induces phenotypic changes, tumor progression and cell invasion. EMBO J. 20, 1910–1920 (2001). A demonstration that a decrease in cancer cell mtDNA content and membrane potential increases cytosolic Ca2+, activates 'retrograde signalling', and increases epithelial–mesenchymal transition and cellular invasiveness.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Biswas, G. et al. A distinctive physiological role for IκBβ in the propagation of mitochondrial respiratory stress signaling. J. Biol. Chem. 283, 12586–12594 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Guha, M., Srinivasan, S., Biswas, G. & Avadhani, N. G. Activation of a novel calcineurin-mediated insulin-like growth factor-1 receptor pathway, altered metabolism, and tumor cell invasion in cells subjected to mitochondrial respiratory stress. J. Biol. Chem. 282, 14536–14546 (2007).

    Article  CAS  PubMed  Google Scholar 

  148. Guha, M., Pan, H., Fang, J. K. & Avadhani, N. G. Heterogeneous nuclear ribonucleoprotein A2 is a common transcriptional coactivator in the nuclear transcription response to mitochondrial respiratory stress. Mol. Biol. Cell 20, 4107–4119 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Guha, M., Fang, J. K., Monks, R., Birnbaum, M. J. & Avadhani, N. G. Activation of Akt is essential for the propagation of mitochondrial respiratory stress signaling and activation of the transcriptional coactivator heterogeneous ribonucleoprotein A2. Mol. Biol. Cell 21, 3578–3589 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Guha, M., Tang, W., Sondheimer, N. & Avadhani, N. G. Role of calcineurin, hnRNPA2 and Akt in mitochondrial respiratory stress-mediated transcription activation of nuclear gene targets. Biochim. Biophys. Acta 1797, 1055–1065 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Fan, W., Lin, C. S., Potluri, P., Procaccio, V. & Wallace, D. C. MtDNA lineage analysis of mouse L cell lines reveals the accumulation of multiple mtDNA mutants and intermolecular recombination. Genes Dev. 26, 384–394 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Park, J. S. et al. A heteroplasmic, not homoplasmic, mitochondrial DNA mutation promotes tumorigenesis via alteration in reactive oxygen species generation and apoptosis. Hum. Mol. Genet. 18, 1578–1589 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Guo, J. Y. et al. Activated Ras requires autophagy to maintain oxidative metabolism and tumorigenesis. Genes Dev. 25, 460–470 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Goh, J. et al. Mitochondrial targeted catalase suppresses invasive breast cancer in mice. BMC Cancer 11, 191 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Woo, D. K. et al. Mitochondrial genome instability and ROS enhance intestinal tumorigenesis in APCMin./+ mice. Am. J. Pathol. 180, 24–31 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Schriner, S. E. et al. Extension of murine life span by overexpression of catalase targeted to mitochondria. Science 308, 1909–1911 (2005).

    Article  CAS  PubMed  Google Scholar 

  157. Zu, X. L. & Guppy, M. Cancer metabolism: facts, fantasy, and fiction. Biochem. Biophys. Res. Commun. 313, 459–465 (2004).

    Article  CAS  PubMed  Google Scholar 

  158. Bonuccelli, G. et al. The reverse Warburg effect: glycolysis inhibitors prevent the tumor promoting effects of caveolin-1 deficient cancer associated fibroblasts. Cell Cycle 9, 1960–1971 (2010).

    Article  CAS  PubMed  Google Scholar 

  159. Pavlides, S. et al. Transcriptional evidence for the “Reverse Warburg Effect” in human breast cancer tumor stroma and metastasis: similarities with oxidative stress, inflammation, Alzheimer's disease, and “Neuron-Glia Metabolic Coupling”. Aging 2, 185–199 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Castello-Cros, R. et al. Matrix remodeling stimulates stromal autophagy, “fueling” cancer cell mitochondrial metabolism and metastasis. Cell Cycle 10, 2021–2034 (2011). This study shows that cancer cell ROS production inactivates stromal cell caveolin 1, thus inducing stromal lactate production that feeds cancer cell oxidative metabolism and growth, a process known as the 'reverse Warburg effect'.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Capparelli, C. et al. Autophagy and senescence in cancer-associated fibroblasts metabolically supports tumor growth and metastasis via glycolysis and ketone production. Cell Cycle 11, 2285–2302 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Pavlides, S. et al. Warburg meets autophagy: cancer-associated fibroblasts accelerate tumor growth and metastasis via oxidative stress, mitophagy, and aerobic glycolysis. Antioxid. Redox Signal. 16, 1264–1284 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Pagliarini, D. J. et al. A mitochondrial protein compendium elucidates complex I disease biology. Cell 134, 112–123 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Phillips, D. et al. Regulation of oxidative phosphorylation complex activity: effects of tissue-specific metabolic stress within an allometric series and acute changes in workload. Am. J. Physiol. Regul. Integr. Comp. Physiol. 302, R1034–R1048 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Wallace, D. C. The epigenome and the mitochondrion: bioenergetics and the environment. Genes Dev. 24, 1571–1573 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Fan, W. et al. A mouse model of mitochondrial disease reveals germline selection against severe mtDNA mutations. Science 319, 958–962 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Hashizume, O. et al. Specific mitochondrial DNA mutation in mice regulates diabetes and lymphoma development. Proc. Natl Acad. Sci. USA 109, 10528–10533 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Oliver, N. A. & Wallace, D. C. Assignment of two mitochondrially synthesized polypeptides to human mitochondrial DNA and their use in the study of intracellular mitochondrial interaction. Mol. Cell. Biol. 2, 30–41 (1982).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The author would like to thank L. Adang and M. Lott for their assistance in preparing this manuscript. This work was supported by the US National Institutes of Health (NIH) grants NS21328, NS070298, AG24373 and DK73691, and a Simons Foundation Grant 205844.

Author information

Authors and Affiliations

Authors

Ethics declarations

Competing interests

The author declares no competing financial interests.

Related links

Related links

DATABASES

National Cancer Institute Drug Dictionary

Pathway Interaction Database

The MITOMAP curated human mitochondrial genome database

Rights and permissions

Reprints and permissions

About this article

Cite this article

Wallace, D. Mitochondria and cancer. Nat Rev Cancer 12, 685–698 (2012). https://doi.org/10.1038/nrc3365

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrc3365

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer