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:

Cancer metabolism: a therapeutic perspective

An Erratum to this article was published on 17 January 2017

This article has been updated

Key Points

  • The metabolic ecology of tumours enables component cells to generate ATP, maintain redox balance, and undertake biosynthesis, which in turn support tumour progression

  • Tumours share features of complex ecosystems, with cancer cells inducing nutrient enrichment; however, the requirement for a tight nutrient balance might be a vulnerability of tumours that can be exploited therapeutically

  • Compared with their normal counterparts, tumour cells require higher rates of catabolite uptake, transfer, and utilization; hence, catabolite-deprivation might be a selective and effective anticancer treatment strategy

  • Targeting glycolysis and mitochondrial metabolism with drug combinations holds promise as another strategy to disrupt the diverse metabolic compartments within tumours

  • Measuring glucose, lactate, pyruvate, β-hydroxybutyrate, and glutamine levels in different tumour compartments and their intercompartmental transfer is needed in clinical trials that examine the efficacy of drugs targeting tumour metabolism

  • Normal tissues frequently have activation of metabolic pathways that are upregulated in cancer and, therefore, dose-limiting toxicity is a challenge in the development of drugs targeting these pathways

Abstract

Awareness that the metabolic phenotype of cells within tumours is heterogeneous — and distinct from that of their normal counterparts — is growing. In general, tumour cells metabolize glucose, lactate, pyruvate, hydroxybutyrate, acetate, glutamine, and fatty acids at much higher rates than their nontumour equivalents; however, the metabolic ecology of tumours is complex because they contain multiple metabolic compartments, which are linked by the transfer of these catabolites. This metabolic variability and flexibility enables tumour cells to generate ATP as an energy source, while maintaining the reduction–oxidation (redox) balance and committing resources to biosynthesis — processes that are essential for cell survival, growth, and proliferation. Importantly, experimental evidence indicates that metabolic coupling between cell populations with different, complementary metabolic profiles can induce cancer progression. Thus, targeting the metabolic differences between tumour and normal cells holds promise as a novel anticancer strategy. In this Review, we discuss how cancer cells reprogramme their metabolism and that of other cells within the tumour microenvironment in order to survive and propagate, thus driving disease progression; in particular, we highlight potential metabolic vulnerabilities that might be targeted therapeutically.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: Metabolic adaptations of cancer cells.
Figure 2: Metabolic heterogeneity in tumours.
Figure 3: Examples of existing and potential anticancer drugs that target metabolic processes.
Figure 4: Metabolic processes regulated by HIF-1 and MYC transcription factors in the catabolic and anabolic regions within a tumour.

Similar content being viewed by others

Change history

  • 17 January 2017

    Owing to a typesetting error, the final line of text in Box 3, and the abbreviation lists for Tables 2 and 3 were omitted from the print and the online pdf versions of this article; for Table 3, the abbrevation list was also omitted from the online html version. These errors have now been corrected in the online pdf and html versions of the manuscript.

References

  1. Ahn, C. S. & Metallo, C. M. Mitochondria as biosynthetic factories for cancer proliferation. Cancer Metab. 3, 1 (2015).

    PubMed  PubMed Central  Google Scholar 

  2. Vander Heiden, M. G., Cantley, L. C. & Thompson, C. B. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science 324, 1029–1033 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Pfeiffer, T., Schuster, S. & Bonhoeffer, S. Cooperation and competition in the evolution of ATP-producing pathways. Science 292, 504–507 (2001).

    CAS  PubMed  Google Scholar 

  4. Cox, E. & Bonner, J. The advantages of togetherness. Science 292, 448–449 (2001).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  6. Martinez-Outschoorn, U. E., Lisanti, M. P. & Sotgia, F. Catabolic cancer-associated fibroblasts transfer energy and biomass to anabolic cancer cells, fueling tumor growth. Semin. Cancer Biol. 25, 47–60 (2014).

    CAS  PubMed  Google Scholar 

  7. Sonveaux, P. et al. Targeting lactate-fueled respiration selectively kills hypoxic tumor cells in mice. J. Clin. Invest. 118, 3930–3942 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Whitaker-Menezes, D. et al. Hyperactivation of oxidative mitochondrial metabolism in epithelial cancer cells in situ: visualizing the therapeutic effects of metformin in tumor tissue. Cell Cycle 10, 4047–4064 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Goodwin, M. L. et al. Modeling alveolar soft part sarcomagenesis in the mouse: a role for lactate in the tumor microenvironment. Cancer Cell 26, 851–862 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Doherty, J. R. & Cleveland, J. L. Targeting lactate metabolism for cancer therapeutics. J. Clin. Invest. 123, 3685–3692 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Pietrocola, F., Galluzzi, L., Bravo- San Pedro, J. M., Madeo, F. & Kroemer, G. Acetyl coenzyme A: a central metabolite and second messenger. Cell Metab. 21, 805–821 (2015).

    CAS  PubMed  Google Scholar 

  12. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell 144, 646–674 (2011).

    CAS  PubMed  Google Scholar 

  13. Viale, A. et al. Oncogene ablation-resistant pancreatic cancer cells depend on mitochondrial function. Nature 514, 628–632 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Wallace, D. C. Mitochondria and cancer Nat. Rev.Cancer 12, 685–698 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Pavlides, S. et al. The reverse Warburg effect: aerobic glycolysis in cancer associated fibroblasts and the tumor stroma. Cell Cycle 8, 3984–4001 (2009).

    CAS  PubMed  Google Scholar 

  16. Nieman, K. M. et al. Adipocytes promote ovarian cancer metastasis and provide energy for rapid tumor growth. Nat. Med. 17, 1498–1503 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Fiaschi, T. et al. Reciprocal metabolic reprogramming through lactate shuttle coordinately influences tumor–stroma interplay. Cancer Res. 72, 5130–5140 (2012).

    CAS  PubMed  Google Scholar 

  18. Boroughs, L. K. & DeBerardinis, R. J. Metabolic pathways promoting cancer cell survival and growth. Nat. Cell Biol. 17, 351–359 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Galluzzi, L., Kepp, O., Vander Heiden, M. G. & Kroemer, G. Metabolic targets for cancer therapy. Nat. Rev. Drug Discov. 12, 829–846 (2013).

    CAS  PubMed  Google Scholar 

  20. Marchiq, I. & Pouyssegur, J. Hypoxia, cancer metabolism and the therapeutic benefit of targeting lactate/H+ symporters. J. Mol. Med. (Berl.) 94, 155–171 (2016).

    CAS  Google Scholar 

  21. Deep, G. & Agarwal, R. Targeting tumor microenvironment with silibinin: promise and potential for a translational cancer chemopreventive strategy. Curr. Cancer Drug Targets 13, 486–499 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Ooi, A. T. & Gomperts, B. N. Molecular pathways: targeting cellular energy metabolism in cancer via inhibition of SLC2A1 and LDHA. Clin. Cancer Res. 21, 2440–2444 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Vander Heiden, M. G. et al. Identification of small molecule inhibitors of pyruvate kinase M2. Biochem. Pharmacol. 79, 1118–1124 (2010).

    CAS  PubMed  Google Scholar 

  24. Cortes-Cros, M. et al. M2 isoform of pyruvate kinase is dispensable for tumor maintenance and growth. Proc. Natl Acad. Sci. USA 110, 489–494 (2013).

    CAS  PubMed  Google Scholar 

  25. Maschek, G. et al. 2-deoxy-d-glucose increases the efficacy of adriamycin and paclitaxel in human osteosarcoma and non-small cell lung cancers in vivo. Cancer Res. 64, 31–34 (2004).

    CAS  PubMed  Google Scholar 

  26. Goldin, N. et al. Methyl jasmonate binds to and detaches mitochondria-bound hexokinase. Oncogene 27, 4636–4643 (2008).

    CAS  PubMed  Google Scholar 

  27. Dwarakanath, B. S. et al. Clinical studies for improving radiotherapy with 2-deoxy-D-glucose: present status and future prospects. J. Cancer Res. Ther. 5 (Suppl. 1), S21–S26 (2009).

    CAS  PubMed  Google Scholar 

  28. Papaldo, P. et al. Addition of either lonidamine or granulocyte colony-stimulating factor does not improve survival in early breast cancer patients treated with high-dose epirubicin and cyclophosphamide. J. Clin. Oncol. 21, 3462–3468 (2003).

    CAS  PubMed  Google Scholar 

  29. DeBerardinis, R. J. & Cheng, T. Q's next: the diverse functions of glutamine in metabolism, cell biology and cancer. Oncogene 29, 313–324 (2009).

    PubMed  PubMed Central  Google Scholar 

  30. Marin-Valencia, I. et al. Analysis of tumor metabolism reveals mitochondrial glucose oxidation in genetically diverse human glioblastomas in the mouse brain in vivo. Cell Metab. 15, 827–837 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Lai, H. S., Lee, J. C., Lee, P. H., Wang, S. T. & Chen, W. J. Plasma free amino acid profile in cancer patients. Semin. Cancer Biol. 15, 267–276 (2005).

    CAS  PubMed  Google Scholar 

  32. Fan, J. et al. Glutamine-driven oxidative phosphorylation is a major ATP source in transformed mammalian cells in both normoxia and hypoxia. Mol. Syst. Biol. 9, 712 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Metallo, C. M. et al. Reductive glutamine metabolism by IDH1 mediates lipogenesis under hypoxia. Nature 481, 380–384 (2011).

    PubMed  PubMed Central  Google Scholar 

  34. Hensley, C. T., Wasti, A. T. & DeBerardinis, R. J. Glutamine and cancer: cell biology, physiology, and clinical opportunities. J. Clin. Invest. 123, 3678–3684 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Mullen, A. R. et al. Reductive carboxylation supports growth in tumour cells with defective mitochondria. Nature 481, 385–388 (2011).

    PubMed  PubMed Central  Google Scholar 

  36. Diehn, M. et al. Association of reactive oxygen species levels and radioresistance in cancer stem cells. Nature 458, 780–783 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Xiang, Y. et al. Targeted inhibition of tumor-specific glutaminase diminishes cell-autonomous tumorigenesis. J. Clin. Invest. 125, 2293–2306 (2015).

    PubMed  PubMed Central  Google Scholar 

  38. Gross, M. I. et al. Antitumor activity of the glutaminase inhibitor CB-839 in triple-negative breast cancer. Mol. Cancer Ther. 13, 890–901 (2014).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Filipp, F. V. et al. Glutamine-fueled mitochondrial metabolism is decoupled from glycolysis in melanoma. Pigment Cell Melanoma Res. 25, 732–739 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Hirayama, A. et al. Quantitative metabolome profiling of colon and stomach cancer microenvironment by capillary electrophoresis time-of-flight mass spectrometry. Cancer Res. 69, 4918–4925 (2009).

    CAS  PubMed  Google Scholar 

  42. Ogawa, T., Washio, J., Takahashi, T., Echigo, S. & Takahashi, N. Glucose and glutamine metabolism in oral squamous cell carcinoma: insight from a quantitative metabolomic approach. Oral Surg. Oral Med. Oral Pathol. Oral Radiol. 118, 218–225 (2014).

    PubMed  Google Scholar 

  43. Hirschhaeuser, F., Sattler, U. G. & Mueller-Klieser, W. Lactate: a metabolic key player in cancer. Cancer Res. 71, 6921–6925 (2011).

    CAS  PubMed  Google Scholar 

  44. Kennedy, K. M. et al. Catabolism of exogenous lactate reveals it as a legitimate metabolic substrate in breast cancer. PLoS ONE 8, e75154 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Walenta, S., Schroeder, T. & Mueller-Klieser, W. Lactate in solid malignant tumors: potential basis of a metabolic classification in clinical oncology. Curr. Med. Chem. 11, 2195–2204 (2004).

    CAS  PubMed  Google Scholar 

  46. Joyce, J. A. & Fearon, D. T. T cell exclusion, immune privilege, and the tumor microenvironment. Science 348, 74–80 (2015).

    CAS  PubMed  Google Scholar 

  47. Colegio, O. R. et al. Functional polarization of tumour-associated macrophages by tumour-derived lactic acid. Nature 513, 559–563 (2013).

    Google Scholar 

  48. Le, A. et al. Inhibition of lactate dehydrogenase A induces oxidative stress and inhibits tumor progression. Proc. Natl Acad. Sci. USA 107, 2037–2042 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Doherty, J. R. et al. Blocking lactate export by inhibiting the Myc target MCT1 disables glycolysis and glutathione synthesis. Cancer Res. 74, 908–920 (2014).

    CAS  PubMed  Google Scholar 

  50. Martinez-Outschoorn, U. E., Sotgia, F. & Lisanti, M. P. Power surge: supporting cells 'fuel' cancer cell mitochondria. Cell Metab. 15, 4–5 (2012).

    CAS  PubMed  Google Scholar 

  51. Sotgia, F. et al. Caveolin-1 and cancer metabolism in the tumor microenvironment: markers, models, and mechanisms. Annu. Rev. Pathol. 7, 423–467 (2012).

    CAS  PubMed  Google Scholar 

  52. Zhang, D. et al. Metabolic reprogramming of cancer-associated fibroblasts by IDH3α downregulation. Cell Rep. 10, 1335–1348 (2015).

    PubMed  Google Scholar 

  53. Polanski, R. et al. Activity of the monocarboxylate transporter 1 inhibitor AZD3965 in small cell lung cancer. Clin. Cancer Res. 20, 926–937 (2014).

    CAS  PubMed  Google Scholar 

  54. Halestrap, A. P. Monocarboxylic acid transport. Compr. Physiol. 3, 1611–1643 (2013).

    PubMed  Google Scholar 

  55. Pellerin, L. & Magistretti, P. J. Sweet sixteen for ANLS. J. Cereb. Blood Flow Metab. 32, 1152–1166 (2012).

    CAS  PubMed  Google Scholar 

  56. Scharfman, H. E. Metabolic control of epilepsy. Science 347, 1312–1313 (2015).

    CAS  PubMed  Google Scholar 

  57. Sada, N., Lee, S., Katsu, T., Otsuki, T. & Inoue, T. Targeting LDH enzymes with a stiripentol analog to treat epilepsy. Science 347, 1362–1367 (2015).

    CAS  PubMed  Google Scholar 

  58. Bailey, K. M., Wojtkowiak, J. W., Hashim, A. I. & Gillies, R. J. Targeting the metabolic microenvironment of tumors. Adv. Pharmacol. 65, 63–107 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Semenza, G. L. Hypoxia-inducible factors: mediators of cancer progression and targets for cancer therapy. Trends Pharmacol. Sci. 33, 207–214 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Rani, R. & Kumar, V. Recent update on human lactate dehydrogenase enzyme 5 (hLDH5) inhibitors: a promising approach for cancer chemotherapy. J. Med. Chem. 59, 487–496 (2016).

    CAS  PubMed  Google Scholar 

  61. Kanno, T. et al. Lactate dehydrogenase M-subunit deficiency: a new type of hereditary exertional myopathy. Clin. Chim. Acta 173, 89–98 (1988).

    CAS  PubMed  Google Scholar 

  62. Diers, A. R., Broniowska, K. A., Chang, C. F. & Hogg, N. Pyruvate fuels mitochondrial respiration and proliferation of breast cancer cells: effect of monocarboxylate transporter inhibition. Biochem. J. 444, 561–571 (2012).

    CAS  PubMed  Google Scholar 

  63. Hatzivassiliou, G. et al. ATP citrate lyase inhibition can suppress tumor cell growth. Cancer Cell 8, 311–321 (2005).

    CAS  PubMed  Google Scholar 

  64. Kamphorst, J. J. et al. Hypoxic and Ras-transformed cells support growth by scavenging unsaturated fatty acids from lysophospholipids. Proc. Natl Acad. Sci. USA 110, 8882–8887 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Gaglio, D. et al. Oncogenic K-Ras decouples glucose and glutamine metabolism to support cancer cell growth. Mol. Syst. Biol. 7, 523 (2011).

    PubMed  PubMed Central  Google Scholar 

  66. Porstmann, T. et al. SREBP activity is regulated by mTORC1 and contributes to Akt-dependent cell growth. Cell Metab. 8, 224–236 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Duvel, K. et al. Activation of a metabolic gene regulatory network downstream of mTOR complex 1. Mol. Cell 39, 171–183 (2010).

    PubMed  PubMed Central  Google Scholar 

  68. Grassian, A. R., Metallo, C. M., Coloff, J. L., Stephanopoulos, G. & Brugge, J. S. Erk regulation of pyruvate dehydrogenase flux through PDK4 modulates cell proliferation. Genes Dev. 25, 1716–1733 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Schug, Z. T. et al. Acetyl-CoA synthetase 2 promotes acetate utilization and maintains cancer cell growth under metabolic stress. Cancer Cell 27, 57–71 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Rabinowitz, J. D. & White, E. Autophagy and metabolism. Science 330, 1344–1348 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Sun, Y. et al. Treatment-induced damage to the tumor microenvironment promotes prostate cancer therapy resistance through WNT16B. Nat. Med. 18, 1359–1368 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Maslowski, K. M. et al. Regulation of inflammatory responses by gut microbiota and chemoattractant receptor GPR43. Nature 461, 1282–1286 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Ley, R. E., Turnbaugh, P. J., Klein, S. & Gordon, J. I. Microbial ecology: human gut microbes associated with obesity. Nature 444, 1022–1023 (2006).

    CAS  PubMed  Google Scholar 

  74. Clemente, J. C., Ursell, L. K., Parfrey, L. W. & Knight, R. The impact of the gut microbiota on human health: an integrative view. Cell 148, 1258–1270 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Abreu, M. T. & Peek, R. M. Jr. Gastrointestinal malignancy and the microbiome. Gastroenterology 146, 1534–1546.e1533 (2014).

    CAS  PubMed  Google Scholar 

  76. Paoli, A., Bosco, G., Camporesi, E. M. & Mangar, D. Ketosis, ketogenic diet and food intake control: a complex relationship. Front. Psychol. 6, 27 (2015).

    PubMed  PubMed Central  Google Scholar 

  77. Kruiswijk, F., Labuschagne, C. F. & Vousden, K. H. p53 in survival, death and metabolic health: a lifeguard with a licence to kill. Nat. Rev. Mol. Cell Biol. 16, 393–405 (2015).

    CAS  PubMed  Google Scholar 

  78. Jeon, S. M., Chandel, N. S. & Hay, N. AMPK regulates NADPH homeostasis to promote tumour cell survival during energy stress. Nature 485, 661–665 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Parks, S. K., Mazure, N. M., Counillon, L. & Pouyssegur, J. Hypoxia promotes tumor cell survival in acidic conditions by preserving ATP levels. J. Cell. Physiol. 228, 1854–1862 (2013).

    CAS  PubMed  Google Scholar 

  80. Caro, P. et al. Metabolic signatures uncover distinct targets in molecular subsets of diffuse large B cell lymphoma. Cancer Cell 22, 547–560 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Woyach, J. A. et al. Resistance mechanisms for the Bruton's tyrosine kinase inhibitor ibrutinib. N. Engl. J. Med. 370, 2286–2294 (2014).

    PubMed  PubMed Central  Google Scholar 

  82. Pike, L. S., Smift, A. L., Croteau, N. J., Ferrick, D. A. & Wu, M. Inhibition of fatty acid oxidation by etomoxir impairs NADPH production and increases reactive oxygen species resulting in ATP depletion and cell death in human glioblastoma cells. Biochim. Biophys. Acta 1807, 726–734 (2011).

    CAS  PubMed  Google Scholar 

  83. Schlaepfer, I. R. et al. Hypoxia induces triglycerides accumulation in prostate cancer cells and extracellular vesicles supporting growth and invasiveness following reoxygenation. Oncotarget 6, 22836–22856 (2015).

    PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  85. Holubarsch, C. J. et al. A double-blind randomized multicentre clinical trial to evaluate the efficacy and safety of two doses of etomoxir in comparison with placebo in patients with moderate congestive heart failure: the ERGO (etomoxir for the recovery of glucose oxidation) study. Clin. Sci. (Lond.) 113, 205–212 (2007).

    CAS  Google Scholar 

  86. Bertolini, F., Sukhatme, V. P. & Bouche, G. Drug repurposing in oncology — patient and health systems opportunities. Nat. Rev. Clin. Oncol. 12, 732–742 (2015).

    PubMed  Google Scholar 

  87. Hamanaka, R. B. & Chandel, N. S. Targeting glucose metabolism for cancer therapy. J. Exp. Med. 209, 211–215 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Wilson, P. M., Danenberg, P. V., Johnston, P. G., Lenz, H. J. & Ladner, R. D. Standing the test of time: targeting thymidylate biosynthesis in cancer therapy. Nat. Rev. Clin. Oncol. 11, 282–298 (2014).

    CAS  PubMed  Google Scholar 

  89. Pieters, R. et al. Pharmacokinetics, pharmacodynamics, efficacy, and safety of a new recombinant asparaginase preparation in children with previously untreated acute lymphoblastic leukemia: a randomized phase 2 clinical trial. Blood 112, 4832–4838 (2008).

    CAS  PubMed  Google Scholar 

  90. Ascierto, P. A. et al. Pegylated arginine deiminase treatment of patients with metastatic melanoma: results from phase I and II studies. J. Clin. Oncol. 23, 7660–7668 (2005).

    CAS  PubMed  Google Scholar 

  91. Glazer, E. S. et al. Phase II study of pegylated arginine deiminase for nonresectable and metastatic hepatocellular carcinoma. J. Clin. Oncol. 28, 2220–2226 (2010).

    CAS  PubMed  Google Scholar 

  92. Zhai, L. et al. Molecular pathways: targeting IDO1 and other tryptophan dioxygenases for cancer immunotherapy. Clin. Cancer Res. 21, 5427–5433 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  93. Possemato, R. et al. Functional genomics reveal that the serine synthesis pathway is essential in breast cancer. Nature 476, 346–350 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Locasale, J. W. et al. Phosphoglycerate dehydrogenase diverts glycolytic flux and contributes to oncogenesis. Nat. Genet. 43, 869–874 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  95. DeNicola, G. M. et al. NRF2 regulates serine biosynthesis in non-small cell lung cancer. Nat. Genet. 47, 1475–1481 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  96. Li, B. & Simon, M. C. Molecular pathways: targeting MYC-induced metabolic reprogramming and oncogenic stress in cancer. Clin. Cancer Res. 19, 5835–5841 (2013).

    CAS  PubMed  Google Scholar 

  97. Stine, Z. E., Walton, Z. E., Altman, B. J., Hsieh, A. L. & Dang, C. V. MYC, metabolism and cancer. Cancer Discov. 5, 1024–1039 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Fan, J. et al. Human phosphoglycerate dehydrogenase produces the oncometabolite d-2-hydroxyglutarate. ACS Chem. Biol. 10, 510–516 (2015).

    CAS  PubMed  Google Scholar 

  99. Agrawal, S., Kumar, A., Srivastava, V. & Mishra, B. N. Cloning, expression, activity and folding studies of serine hydroxymethyltransferase: a target enzyme for cancer chemotherapy. J. Mol. Microbiol. Biotechnol. 6, 67–75 (2003).

    CAS  PubMed  Google Scholar 

  100. Fan, J. et al. Quantitative flux analysis reveals folate-dependent NADPH production. Nature 510, 298–302 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  101. Jain, M. et al. Metabolite profiling identifies a key role for glycine in rapid cancer cell proliferation. Science 336, 1040–1044 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  102. Bauer, D. E., Hatzivassiliou, G., Zhao, F., Andreadis, C. & Thompson, C. B. ATP citrate lyase is an important component of cell growth and transformation. Oncogene 24, 6314–6322 (2005).

    CAS  PubMed  Google Scholar 

  103. Chajes, V., Cambot, M., Moreau, K., Lenoir, G. M. & Joulin, V. Acetyl-CoA carboxylase α is essential to breast cancer cell survival. Cancer Res. 66, 5287–5294 (2006).

    CAS  PubMed  Google Scholar 

  104. Clem, B. F. et al. A novel small molecule antagonist of choline kinase-alpha that simultaneously suppresses MAPK and PI3K/AKT signaling. Oncogene 30, 3370–3380 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  105. Flavin, R., Peluso, S., Nguyen, P. L. & Loda, M. Fatty acid synthase as a potential therapeutic target in cancer. Future Oncol. 6, 551–562 (2010).

    CAS  PubMed  Google Scholar 

  106. Mulvihill, M. M. & Nomura, D. K. Therapeutic potential of monoacylglycerol lipase inhibitors. Life Sci. 92, 492–497 (2013).

    CAS  PubMed  Google Scholar 

  107. Gallego-Ortega, D., Gomez del Pulgar, T., Valdes-Mora, F., Cebrian, A. & Lacal, J. C. Involvement of human choline kinase alpha and beta in carcinogenesis: a different role in lipid metabolism and biological functions. Adv. Enzyme Regul. 51, 183–194 (2011).

    CAS  PubMed  Google Scholar 

  108. Glunde, K., Bhujwalla, Z. M. & Ronen, S. M. Choline metabolism in malignant transformation. Nat. Rev. Cancer 11, 835–848 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Kubatka, P., Kruzliak, P., Rotrekl, V., Jelinkova, S. & Mladosievicova, B. Statins in oncological research: from experimental studies to clinical practice. Crit. Rev. Oncol. Hematol. 92, 296–311 (2014).

    PubMed  Google Scholar 

  110. Nielsen, S. F., Nordestgaard, B. G. & Bojesen, S. E. Statin use and reduced cancer-related mortality. N. Engl. J. Med. 367, 1792–1802 (2012).

    CAS  PubMed  Google Scholar 

  111. Sriskanthadevan, S. et al. AML cells have low spare reserve capacity in their respiratory chain that renders them susceptible to oxidative metabolic stress. Blood 125, 2120–2130 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Funes, J. M. et al. Transformation of human mesenchymal stem cells increases their dependency on oxidative phosphorylation for energy production. Proc. Natl Acad. Sci. USA 104, 6223–6228 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Fogal, V. et al. Mitochondrial p32 protein is a critical regulator of tumor metabolism via maintenance of oxidative phosphorylation. Mol. Cell. Biol. 30, 1303–1318 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  114. Curry, J. M. et al. Cancer metabolism, stemness and tumor recurrence: MCT1 and MCT4 are functional biomarkers of metabolic symbiosis in head and neck cancer. Cell Cycle 12, 1371–1384 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  115. Wurm, C. A. et al. Nanoscale distribution of mitochondrial import receptor Tom20 is adjusted to cellular conditions and exhibits an inner-cellular gradient. Proc. Natl Acad. Sci. USA 108, 13546–13551 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Gehrke, S. et al. PINK1 and Parkin control localized translation of respiratory chain component mRNAs on mitochondria outer membrane. Cell Metab. 21, 95–108 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  117. Vazquez, F. et al. PGC1α expression defines a subset of human melanoma tumors with increased mitochondrial capacity and resistance to oxidative stress. Cancer Cell 23, 287–301 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  118. LeBleu, V. S. et al. PGC-1α mediates mitochondrial biogenesis and oxidative phosphorylation in cancer cells to promote metastasis. Nat. Cell Biol. 16, 992–1003 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  119. Tan, A. S. et al. Mitochondrial genome acquisition restores respiratory function and tumorigenic potential of cancer cells without mitochondrial DNA. Cell Metab. 21, 81–94 (2015).

    CAS  PubMed  Google Scholar 

  120. Katajisto, P. et al. Asymmetric apportioning of aged mitochondria between daughter cells is required for stemness. Science 348, 340–343 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  121. Lamb, R. et al. Antibiotics that target mitochondria effectively eradicate cancer stem cells, across multiple tumor types: treating cancer like an infectious disease. Oncotarget 6, 4569–4584 (2015).

    PubMed  PubMed Central  Google Scholar 

  122. Lamb, R. et al. Dissecting tumor metabolic heterogeneity: telomerase and large cell size metabolically define a sub-population of stem-like, mitochondrial-rich, cancer cells. Oncotarget 6, 21892–21905 (2015).

    PubMed  PubMed Central  Google Scholar 

  123. Lo-Coco, F. et al. Retinoic acid and arsenic trioxide for acute promyelocytic leukemia. N. Engl. J. Med. 369, 111–121 (2013).

    CAS  PubMed  Google Scholar 

  124. Owen, M. R., Doran, E. & Halestrap, A. P. Evidence that metformin exerts its anti-diabetic effects through inhibition of complex 1 of the mitochondrial respiratory chain. Biochem. J. 348, 607–614 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  125. Pelicano, H. et al. Inhibition of mitochondrial respiration: a novel strategy to enhance drug-induced apoptosis in human leukemia cells by a reactive oxygen species-mediated mechanism. J. Biol. Chem. 278, 37832–37839 (2003).

    CAS  PubMed  Google Scholar 

  126. Pollak, M. Overcoming drug development bottlenecks with repurposing: repurposing biguanides to target energy metabolism for cancer treatment. Nat. Med. 20, 591–593 (2014).

    CAS  PubMed  Google Scholar 

  127. Bowker, S. L., Majumdar, S. R., Veugelers, P. & Johnson, J. A. Increased cancer-related mortality for patients with type 2 diabetes who use sulfonylureas or insulin. Diabetes Care 29, 254–258 (2006).

    PubMed  Google Scholar 

  128. Hirsch, H. A., Iliopoulos, D., Tsichlis, P. N. & Struhl, K. Metformin selectively targets cancer stem cells, and acts together with chemotherapy to block tumor growth and prolong remission. Cancer Res. 69, 7507–7511 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  129. Hadad, S. et al. Evidence for biological effects of metformin in operable breast cancer: a pre-operative, window-of-opportunity, randomized trial. Breast Cancer Res. Treat. 128, 783–794 (2011).

    CAS  PubMed  Google Scholar 

  130. Menendez, J. A. et al. Metformin is synthetically lethal with glucose withdrawal in cancer cells. Cell Cycle 11, 2782–2792 (2012).

    CAS  PubMed  Google Scholar 

  131. Birsoy, K. et al. Metabolic determinants of cancer cell sensitivity to glucose limitation and biguanides. Nature 508, 108–112 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  132. Chen, G., Xu, S., Renko, K. & Derwahl, M. Metformin inhibits growth of thyroid carcinoma cells, suppresses self-renewal of derived cancer stem cells, and potentiates the effect of chemotherapeutic agents. J. Clin. Endocrinol. Metab. 97, E510–E520 (2012).

    CAS  PubMed  Google Scholar 

  133. Song, C. W. et al. Metformin kills and radiosensitizes cancer cells and preferentially kills cancer stem cells. Sci. Rep. 2, 362 (2012).

    PubMed  PubMed Central  Google Scholar 

  134. Bao, B. et al. Metformin inhibits cell proliferation, migration and invasion by attenuating CSC function mediated by deregulating miRNAs in pancreatic cancer cells. Cancer Prev. Res. (Phila.) 5, 355–364 (2012).

    CAS  Google Scholar 

  135. Foretz, M., Guigas, B., Bertrand, L., Pollak, M. & Viollet, B. Metformin: from mechanisms of action to therapies. Cell Metab. 20, 953–966 (2014).

    CAS  PubMed  Google Scholar 

  136. Kordes, S. et al. Metformin in patients with advanced pancreatic cancer: a double-blind, randomised, placebo-controlled phase 2 trial. Lancet Oncol. 16, 839–847 (2015).

    CAS  PubMed  Google Scholar 

  137. Salani, B. et al. Metformin, cancer and glucose metabolism. Endocr. Relat. Cancer 21, R461–R471 (2014).

    PubMed  Google Scholar 

  138. Britten, C. D. et al. A phase I and pharmacokinetic study of the mitochondrial-specific rhodacyanine dye analog MKT 077. Clin. Cancer Res. 6, 42–49 (2000).

    CAS  PubMed  Google Scholar 

  139. Propper, D. J. et al. Phase I trial of the selective mitochondrial toxin MKT077 in chemo-resistant solid tumours. Ann. Oncol. 10, 923–927 (1999).

    CAS  PubMed  Google Scholar 

  140. Michelakis, E. D., Webster, L. & Mackey, J. R. Dichloroacetate (DCA) as a potential metabolic-targeting therapy for cancer. Br. J. Cancer 99, 989–994 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  141. Fu, X. et al. 2-hydroxyglutarate inhibits ATP synthase and mTOR signaling. Cell Metab. 22, 508–515 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  142. Yan, H., Bigner, D. D., Velculescu, V. & Parsons, D. W. Mutant metabolic enzymes are at the origin of gliomas. Cancer Res. 69, 9157–9159 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  143. Fathi, A. T., Wander, S. A., Faramand, R. & Emadi, A. Biochemical, epigenetic and metabolic approaches to target IDH mutations in acute myeloid leukemia. Semin Hematol. 52, 165–171 (2015).

    CAS  PubMed  Google Scholar 

  144. Sagan, L. On the origin of mitosing cells. J. Theor. Biol. 14, 255–274 (1967).

    CAS  PubMed  Google Scholar 

  145. Beckmann, R. & Herrmann, J. M. Mitoribosome oddities. Science 348, 288–289 (2015).

    CAS  PubMed  Google Scholar 

  146. Greber, B. J. et al. The complete structure of the 55S mammalian mitochondrial ribosome. Science 348, 303–308 (2015).

    CAS  PubMed  Google Scholar 

  147. Amunts, A., Brown, A., Toots, J., Scheres, S. H. & Ramakrishnan, V. The structure of the human mitochondrial ribosome. Science 348, 95–98 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  148. Amunts, A. et al. Structure of the yeast mitochondrial large ribosomal subunit. Science 343, 1485–1489 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  149. Brown, A. et al. Structure of the large ribosomal subunit from human mitochondria. Science 346, 718–722 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  150. Wilson, D. N. Ribosome-targeting antibiotics and mechanisms of bacterial resistance. Nat. Rev. Microbiol. 12, 35–48 (2014).

    CAS  PubMed  Google Scholar 

  151. Carter, A. P. et al. Functional insights from the structure of the 30S ribosomal subunit and its interactions with antibiotics. Nature 407, 340–348 (2000).

    CAS  PubMed  Google Scholar 

  152. Moazed, D. & Noller, H. F. Interaction of antibiotics with functional sites in 16S ribosomal RNA. Nature 327, 389–394 (1987).

    CAS  PubMed  Google Scholar 

  153. Greber, B. J. & Ban, N. Structure and function of the mitochondrial ribosome. Annu. Rev. Biochem. http://dx.doi.org/10.1146/annurevbiochem060815014343 (2016).

  154. Prezant, T. R. et al. Mitochondrial ribosomal RNA mutation associated with both antibiotic-induced and non-syndromic deafness. Nat. Genet. 4, 289–294 (1993).

    CAS  PubMed  Google Scholar 

  155. Bitner-Glindzicz, M. et al. Prevalence of mitochondrial 1555A→G mutation in European children. N. Engl. J. Med. 360, 640–642 (2009).

    PubMed  Google Scholar 

  156. Soriano, A., Miro, O. & Mensa, J. Mitochondrial toxicity associated with linezolid. N. Engl. J. Med. 353, 2305–2306 (2005).

    CAS  PubMed  Google Scholar 

  157. Ferreri, A. J. et al. Bacteria-eradicating therapy with doxycycline in ocular adnexal MALT lymphoma: a multicenter prospective trial. J. Natl Cancer Inst. 98, 1375–1382 (2006).

    CAS  PubMed  Google Scholar 

  158. Ferreri, A. J. et al. Chlamydophila psittaci eradication with doxycycline as first-line targeted therapy for ocular adnexae lymphoma: final results of an international phase II trial. J. Clin. Oncol. 30, 2988–2994 (2012).

    CAS  PubMed  Google Scholar 

  159. Skrtic, M. et al. Inhibition of mitochondrial translation as a therapeutic strategy for human acute myeloid leukemia. Cancer Cell 20, 674–688 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  160. Bahrami, F., Morris, D. L. & Pourgholami, M. H. Tetracyclines: drugs with huge therapeutic potential. Mini Rev. Med. Chem. 12, 44–52 (2012).

    CAS  PubMed  Google Scholar 

  161. Yang, M., Soga, T. & Pollard, P. J. Oncometabolites: linking altered metabolism with cancer. J. Clin. Invest. 123, 3652–3658 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  162. Vazquez, A., Liu, J., Zhou, Y. & Oltvai, Z. N. Catabolic efficiency of aerobic glycolysis: the Warburg effect revisited. BMC Syst. Biol. 4, 58 (2010).

    PubMed  PubMed Central  Google Scholar 

  163. Shlomi, T., Benyamini, T., Gottlieb, E., Sharan, R. & Ruppin, E. Genome-scale metabolic modeling elucidates the role of proliferative adaptation in causing the Warburg effect. PLoS Comput. Biol. 7, e1002018 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  165. Martinez-Outschoorn, U. E. et al. Oxidative stress in cancer associated fibroblasts drives tumor-stroma co-evolution: a new paradigm for understanding tumor metabolism, the field effect and genomic instability in cancer cells. Cell Cycle 9, 3256–3276 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  166. Asterholm, I. W., Mundy, D. I., Weng, J., Anderson, R. G. & Scherer, P. E. Altered mitochondrial function and metabolic inflexibility associated with loss of caveolin-1. Cell Metab. 15, 171–185 (2012).

    PubMed  PubMed Central  Google Scholar 

  167. Chen, J. L. et al. The genomic analysis of lactic acidosis and acidosis response in human cancers. PLoS Genet. 4, e1000293 (2008).

    PubMed  PubMed Central  Google Scholar 

  168. Xie, J. et al. Beyond Warburg effect — dual metabolic nature of cancer cells. Sci. Rep. 4, 4927 (2014).

    PubMed  PubMed Central  Google Scholar 

  169. Gorrini, C., Harris, I. S. & Mak, T. W. Modulation of oxidative stress as an anticancer strategy. Nat. Rev. Drug Discov. 12, 931–947 (2013).

    CAS  PubMed  Google Scholar 

  170. Nisoli, E. et al. Mitochondrial biogenesis in mammals: the role of endogenous nitric oxide. Science 299, 896–899 (2003).

    CAS  PubMed  Google Scholar 

  171. Nisoli, E. et al. Calorie restriction promotes mitochondrial biogenesis by inducing the expression of eNOS. Science 310, 314–317 (2005).

    CAS  PubMed  Google Scholar 

  172. Hasmann, M. & Schemainda, I. FK866, a highly specific noncompetitive inhibitor of nicotinamide phosphoribosyltransferase, represents a novel mechanism for induction of tumor cell apoptosis. Cancer Res. 63, 7436–7442 (2003).

    CAS  PubMed  Google Scholar 

  173. Sampath, D., Zabka, T. S., Misner, D. L., O'Brien, T. & Dragovich, P. S. Inhibition of nicotinamide phosphoribosyltransferase (NAMPT) as a therapeutic strategy in cancer. Pharmacol. Ther. 151, 16–31 (2015).

    CAS  PubMed  Google Scholar 

  174. Burgos, E. S. NAMPT in regulated NAD biosynthesis and its pivotal role in human metabolism. Curr. Med. Chem. 18, 1947–1961 (2011).

    CAS  PubMed  Google Scholar 

  175. Yun, J. et al. Vitamin C selectively kills KRAS and BRAF mutant colorectal cancer cells by targeting GAPDH. Science 350, 1391–1396 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  176. Porporato, P. E. et al. A mitochondrial switch promotes tumor metastasis. Cell Rep. 8, 754–766 (2014).

    CAS  PubMed  Google Scholar 

  177. DeNicola, G. M. et al. Oncogene-induced Nrf2 transcription promotes ROS detoxification and tumorigenesis. Nature 475, 106–109 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  178. Wells, G. Peptide and small molecule inhibitors of the Keap1–Nrf2 protein–protein interaction. Biochem. Soc. Trans. 43, 674–679 (2015).

    CAS  PubMed  Google Scholar 

  179. Wang, G. L., Jiang, B. H., Rue, E. A. & Semenza, G. L. Hypoxia-inducible factor 1 is a basic-helix-loop- helix-PAS heterodimer regulated by cellular O2 tension. Proc. Natl Acad. Sci. USA 92, 5510–5514 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  180. 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).

    CAS  PubMed  Google Scholar 

  181. Kim, J. W., Gao, P., Liu, Y. C., Semenza, G. L. & Dang, C. V. Hypoxia-inducible factor 1 and dysregulated c-Myc cooperatively induce vascular endothelial growth factor and metabolic switches hexokinase 2 and pyruvate dehydrogenase kinase 1. Mol. Cell. Biol. 27, 7381–7393 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  185. Semenza, G. L. HIF-1 mediates metabolic responses to intratumoral hypoxia and oncogenic mutations. J. Clin. Invest. 123, 3664–3671 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  186. Velasco-Hernandez, T., Hyrenius-Wittsten, A., Rehn, M., Bryder, D. & Cammenga, J. HIF-1α can act as a tumor suppressor gene in murine acute myeloid leukemia. Blood 124, 3597–3607 (2014).

    CAS  PubMed  Google Scholar 

  187. Chiavarina, B. et al. HIF1-alpha functions as a tumor promoter in cancer associated fibroblasts, and as a tumor suppressor in breast cancer cells: autophagy drives compartment-specific oncogenesis. Cell Cycle 9, 3534–3551 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  188. Wilson, W. R. & Hay, M. P. Targeting hypoxia in cancer therapy. Nat. Rev. Cancer 11, 393–410 (2011).

    CAS  PubMed  Google Scholar 

  189. Talekar, M., Boreddy, S. R., Singh, A. & Amiji, M. Tumor aerobic glycolysis: new insights into therapeutic strategies with targeted delivery. Expert Opin. Biol. Ther. 14, 1145–1159 (2014).

    CAS  PubMed  Google Scholar 

  190. Chen, B. J., Wu, Y. L., Tanaka, Y. & Zhang, W. Small molecules targeting c-Myc oncogene: promising anti-cancer therapeutics. Int. J. Biol. Sci. 10, 1084–1096 (2014).

    PubMed  PubMed Central  Google Scholar 

  191. Yuneva, M. O. et al. The metabolic profile of tumors depends on both the responsible genetic lesion and tissue type. Cell Metab. 15, 157–170 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  192. Valencia, T. et al. Metabolic reprogramming of stromal fibroblasts through p62–mTORC1 signaling promotes inflammation and tumorigenesis. Cancer Cell 26, 121–135 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  193. Amend, S. R. & Pienta, K. J. Ecology meets cancer biology: the cancer swamp promotes the lethal cancer phenotype. Oncotarget 6, 9669–9678 (2015).

    PubMed  PubMed Central  Google Scholar 

  194. Draoui, N. & Feron, O. Lactate shuttles at a glance: from physiological paradigms to anti-cancer treatments. Dis. Model. Mech. 4, 727–732 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  195. White, E. Deconvoluting the context-dependent role for autophagy in cancer. Nat. Rev. Cancer 12, 401–410 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  196. Martinez-Outschoorn, U. E. et al. Autophagy in cancer associated fibroblasts promotes tumor cell survival: role of hypoxia, HIF1 induction and NFκB activation in the tumor stromal microenvironment. Cell Cycle 9, 3515–3533 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  197. Takeuchi, H. et al. Synergistic augmentation of rapamycin-induced autophagy in malignant glioma cells by phosphatidylinositol 3-kinase/protein kinase B inhibitors. Cancer Res. 65, 3336–3346 (2005).

    CAS  PubMed  Google Scholar 

  198. Chiarini, F., Evangelisti, C., McCubrey, J. A. & Martelli, A. M. Current treatment strategies for inhibiting mTOR in cancer. Trends Pharmacol. Sci. 36, 124–135 (2015).

    CAS  PubMed  Google Scholar 

  199. Kimura, T., Takabatake, Y., Takahashi, A. & Isaka, Y. Chloroquine in cancer therapy: a double-edged sword of autophagy. Cancer Res. 73, 3–7 (2013).

    CAS  PubMed  Google Scholar 

  200. Farber, S. & Diamond, L. K. Temporary remissions in acute leukemia in children produced by folic acid antagonist, 4-aminopteroyl-glutamic acid. N. Engl. J. Med. 238, 787–793 (1948).

    CAS  PubMed  Google Scholar 

  201. Visentin, M., Zhao, R. & Goldman, I. D. The antifolates. Hematol. Oncol. Clin. North Am. 26, 629–648 (2012).

    PubMed  PubMed Central  Google Scholar 

  202. Kaelin, W. G. Jr & McKnight, S. L. Influence of metabolism on epigenetics and disease. Cell 153, 56–69 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  203. Ye, D., Ma, S., Xiong, Y. & Guan, K. L. R-2-hydroxyglutarate as the key effector of IDH mutations promoting oncogenesis. Cancer Cell 23, 274–276 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  204. Dang, L. et al. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature 462, 739–744 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  205. Losman, J. A. et al. (R)-2-hydroxyglutarate is sufficient to promote leukemogenesis and its effects are reversible. Science 339, 1621–1625 (2013).

    CAS  PubMed  Google Scholar 

  206. Rohle, D. et al. An inhibitor of mutant IDH1 delays growth and promotes differentiation of glioma cells. Science 340, 626–630 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  207. Smolkova, K., Dvorak, A., Zelenka, J., Vitek, L. & Jezek, P. Reductive carboxylation and 2-hydroxyglutarate formation by wild-type IDH2 in breast carcinoma cells. Int. J. Biochem. Cell Biol. 65, 125–133 (2015).

    CAS  PubMed  Google Scholar 

  208. Intlekofer, A. M. et al. Hypoxia induces production of L-2-hydroxyglutarate. Cell Metab. 22, 304–311 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  209. Carrer, A. & Wellen, K. E. Metabolism and epigenetics: a link cancer cells exploit. Curr. Opin. Biotechnol. 34, 23–29 (2015).

    CAS  PubMed  Google Scholar 

  210. Peiris-Pages, M., Martinez-Outschoorn, U. E., Sotgia, F. & Lisanti, M. P. Metastasis and oxidative stress: are antioxidants a metabolic driver of progression? Cell Metab. 22, 956–958 (2015).

    CAS  PubMed  Google Scholar 

  211. Clem, B. F. et al. Targeting 6-phosphofructo-2-kinase (PFKFB3) as a therapeutic strategy against cancer. Mol. Cancer Ther. 12, 1461–1470 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  212. Morais-Santos, F. et al. Targeting lactate transport suppresses in vivo breast tumour growth. Oncotarget 6, 19177–19189 (2015).

    PubMed  PubMed Central  Google Scholar 

  213. Fujiwara, S. et al. PDK1 inhibition is a novel therapeutic target in multiple myeloma. Br. J. Cancer 108, 170–178 (2013).

    CAS  PubMed  Google Scholar 

  214. Sellers, K. et al. Pyruvate carboxylase is critical for non-small-cell lung cancer proliferation. J. Clin. Invest. 125, 687–698 (2015).

    PubMed  PubMed Central  Google Scholar 

  215. Pardee, T. S. et al. A phase I study of the first-in-class antimitochondrial metabolism agent, CPI-613, in patients with advanced hematologic malignancies. Clin. Cancer Res. 20, 5255–5264 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  216. Wang, F. et al. Targeted inhibition of mutant IDH2 in leukemia cells induces cellular differentiation. Science 340, 622–626 (2013).

    CAS  PubMed  Google Scholar 

  217. El-Mir, M. Y. et al. Dimethylbiguanide inhibits cell respiration via an indirect effect targeted on the respiratory chain complex I. J. Biol. Chem. 275, 223–228 (2000).

    CAS  PubMed  Google Scholar 

  218. Jara, J. A. & Lopez-Munoz, R. Metformin and cancer: between the bioenergetic disturbances and the antifolate activity. Pharmacol. Res. 101, 102–108 (2015).

    CAS  PubMed  Google Scholar 

  219. Hitosugi, T. et al. Phosphoglycerate mutase 1 coordinates glycolysis and biosynthesis to promote tumor growth. Cancer Cell 22, 585–600 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  220. Feun, L. G., Kuo, M. T. & Savaraj, N. Arginine deprivation in cancer therapy. Curr. Opin. Clin. Nutr. Metab. Care 18, 78–82 (2015).

    CAS  PubMed  Google Scholar 

  221. Lamb, R. et al. Doxycycline down-regulates DNA-PK and radiosensitizes tumor initiating cells: implications for more effective radiation therapy. Oncotarget 6, 14005–14025 (2015).

    PubMed  PubMed Central  Google Scholar 

  222. Pham, E. et al. Translational impact of nanoparticle-drug conjugate CRLX101 with or without bevacizumab in advanced ovarian cancer. Clin. Cancer Res. 21, 808–818 (2015).

    CAS  PubMed  Google Scholar 

  223. Galluzzi, L. et al. Autophagy in malignant transformation and cancer progression. EMBO J. 34, 856–880 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  224. Zadra, G., Batista, J. L. & Loda, M. Dissecting the dual role of AMPK in cancer: from experimental to human studies. Mol. Cancer Res. 13, 1059–1072 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  225. Sayin, V. I. et al. Antioxidants accelerate lung cancer progression in mice. Sci. Transl. Med. 6, 221ra215 (2014).

    Google Scholar 

  226. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The Sotgia and Lisanti laboratories in the UK have been supported, in part, by funding from the EU (European Research Council Advanced Grant), Breast Cancer Now, The Healthy Life Foundation, and the Manchester Cancer Research Centre (MCRC). The work of Ubaldo E. Martinez-Outschoorn has been supported by the National Cancer Institute (NCI) of the National Institutes of Health (NIH), under Award Number K08-CA175193. Richard G. Pestell's laboratory receives funding from the NIH and the NCI, as well as the Breast Cancer Research Foundation and the Ralph and Marian C. Falk Medical Research Trust.

Author information

Authors and Affiliations

Authors

Contributions

U.E.M.-O. and M.P.-P. researched the data for the article and wrote the manuscript. U.E.M.-O., M.P.-P., F.S., and M.P.L. contributed substantially to discussions of content, and U.E.M.-O., R.G.P., F.S., and M.P.L. reviewed and edited the manuscript before submission.

Corresponding authors

Correspondence to Federica Sotgia or Michael P. Lisanti.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Martinez-Outschoorn, U., Peiris-Pagés, M., Pestell, R. et al. Cancer metabolism: a therapeutic perspective. Nat Rev Clin Oncol 14, 11–31 (2017). https://doi.org/10.1038/nrclinonc.2016.60

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrclinonc.2016.60

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