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.

Why do cancers have high aerobic glycolysis?

Key Points

  • Widespread clinical use of 18fluorodeoxyglucose positron-emission tomography has demonstrated that the glycolytic phenotype is observed in most human cancers.

  • The concept of carcinogenesis as a process that occurs by somatic evolution clearly implies that common traits of the malignant phenotype, such as upregulation of glycolysis, are the result of active selection processes and must confer a significant, identifiable growth advantage.

  • Constitutive upregulation of glycolysis is likely to be an adaptation to hypoxia that develops as pre-malignant lesions grow progressively further from their blood supply. At this stage, the blood supply remains physically separated from the growing cells by an intact basement membrane.

  • Increased acid production from upregulation of glycolysis results in microenvironmental acidosis and requires further adaptation through somatic evolution to phenotypes resistant to acid-induced toxicity.

  • Cell populations that emerge from this evolutionary sequence have a powerful growth advantage, as they alter their environment through increased glycolysis in a way that is toxic to other phenotypes, but harmless to themselves. The environmental acidosis also facilitates invasion through destruction of adjacent normal populations, degradation of the extracellular matrix and promotion of angiogenesis.

  • We propose that the glycolytic phenotype, by conferring a powerful growth advantage, is necessary for evolution of invasive human cancers.

Abstract

If carcinogenesis occurs by somatic evolution, then common components of the cancer phenotype result from active selection and must, therefore, confer a significant growth advantage. A near-universal property of primary and metastatic cancers is upregulation of glycolysis, resulting in increased glucose consumption, which can be observed with clinical tumour imaging. We propose that persistent metabolism of glucose to lactate even in aerobic conditions is an adaptation to intermittent hypoxia in pre-malignant lesions. However, upregulation of glycolysis leads to microenvironmental acidosis requiring evolution to phenotypes resistant to acid-induced cell toxicity. Subsequent cell populations with upregulated glycolysis and acid resistance have a powerful growth advantage, which promotes unconstrained proliferation and invasion.

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: Glucose metabolism in mammalian cells.
Figure 2: Positron-emission tomography imaging with 18fluorodeoxyglucose of a patient with lymphoma.
Figure 3: Pasteur and Warburg effects in non-invasive and metastatic breast cancer cell lines.
Figure 4: Hyperacidity of tumours.
Figure 5: Late-stage ductal carcinoma in situ.
Figure 6: Model for cell–environment interactions in carcinogenesis.

References

  1. Bernards, R. & Weinberg, R. A. A progression puzzle. Nature 418, 823 (2002). A compelling opinion piece, in which the authors convincingly argue that the molecular phenotypes of metastatic cancers arose early during carcinogenesis. Although somatic evolution is implied in this work, the environmental nature of the selection pressures are not discussed.

    CAS  Google Scholar 

  2. Racker, E. History of the Pasteur effect and its pathobiology. Mol. Cell. Biochem. 5, 17–23 (1974).

    CAS  PubMed  Google Scholar 

  3. Warburg, O. Ueber den stoffwechsel der tumoren. (Constable, London, 1930).

    Google Scholar 

  4. Semenza, G. L. et al. 'The metabolism of tumours': 70 years later. Novartis Found. Symp. 240, 251–260 (2001). In this timely review, Semenza describes the relation between HIF1α and the regulation of glycolysis.

    CAS  PubMed  Google Scholar 

  5. Weinhouse, S. The Warburg hypothesis fifty years later. Z. Krebsforsch. Klin. Onkol. Cancer Res. Clin. Oncol. 87, 115–126 (1976).

    CAS  PubMed  Google Scholar 

  6. Hawkins, R. A. & Phelps, M. E. PET in clinical oncology. Cancer Metastasis Rev. 7, 119–142 (1988).

    CAS  PubMed  Google Scholar 

  7. Weber, W. A., Avril, N. & Schwaiger, M. Relevance of positron emission tomography (PET) in oncology. Strahlenther. Onkol. 175, 356–373 (1999).

    CAS  PubMed  Google Scholar 

  8. Gambhir, S. S. Molecular imaging of cancer with positron emission tomography. Nature Rev. Cancer 2, 683–693 (2002). A well written review on FdG PET imaging.

    CAS  Google Scholar 

  9. Czernin, J. & Phelps, M. E. Positron emission tomography scanning: current and future applications. Annu. Rev. Med. 53, 89–112 (2002). A comprehensive review of extant literature. The authors convincingly document the very high sensitivity and specificity of FdG PET in diagnosing and staging diverse types of metastatic cancers.

    CAS  PubMed  Google Scholar 

  10. Bos, R. et al. Biologic correlates of 18fluorodeoxyglucose uptake in human breast cancer measured by positron emission tomography. J.Clin.Oncol. 20, 379–387 (2002). This well-conducted study quantitatively analysed the molecular phenotypes of tumours that had either high or low rates of FdG trapping.

    CAS  PubMed  Google Scholar 

  11. Burt, B. M. et al. Using positron emission tomography with [18F]FDG to predict tumor behavior in experimental colorectal cancer. Neoplasia (New York) 3, 189–195 (2001).

    CAS  Google Scholar 

  12. Schilling, C. H., Schuster, S., Palsson, B. O. & Heinrich, R. Metabolic pathway analysis: basic concepts and scientific applications in the post-genomic era. Biotechnol. Prog. 15, 296–303 (1999).

    CAS  PubMed  Google Scholar 

  13. Dang, C. V., Lewis, B. C., Dolde, C., Dang, G. & Shim, H. Oncogenes in tumor metabolism, tumorigenesis, and apoptosis. J. Bioenerg. Biomembr. 29, 345–354 (1997). One of many papers in this issue of the Journal of Bioenergetics and Biomembranes that dealt with the molecular controls of glucose metabolism. In this review, primary data were presented to support the importance and molecular controls of the glucose transporter and its regulation by MYC.

    CAS  PubMed  Google Scholar 

  14. Rivenzon-Segal, D., Boldin-Adamsky, S., Seger, D., Seger, R. & Degani, H. Glycolysis and glucose transporter 1 as markers of response to hormonal therapy in breast cancer. Int. J. Cancer 107, 177–182 (2003). One of many papers that demonstrates the important role of the glucose transporter in regulating glycolytic flux.

    CAS  PubMed  Google Scholar 

  15. Artemov, D., Bhujwalla, Z. M., Pilatus, U. & Glickson, J. D. Two-compartment model for determination of glycolytic rates of solid tumors by in vivo13C NMR spectroscopy. NMR Biomed. 11, 395–404 (1998).

    CAS  PubMed  Google Scholar 

  16. Mathupala, S. P., Rempel, A. & Pedersen, P. L. Aberrant glycolytic metabolism of cancer cells: a remarkable coordination of genetic, transcriptional, post-translational, and mutational events that lead to a critical role for type II hexokinase. J. Bioenerg. Biomembr. 29, 339–343 (1997). Provides a cogent argument for the role of hexokinase in regulating glycolytic flux and its regulation by oncogenes and subcellular localization.

    CAS  PubMed  Google Scholar 

  17. Kunkel, M. et al. Overexpression of Glut-1 and increased glucose metabolism in tumors are associated with a poor prognosis in patients with oral squamous cell carcinoma. Cancer 97, 1015–1024 (2003). This careful study is one of many that document the diagnostic importance of GLUT1 and glycolysis in carcinomas.

    CAS  PubMed  Google Scholar 

  18. Mochiki, E. et al. Evaluation of 18F-2-deoxy-2-fluoro-D-glucose positron emission tomography for gastric cancer. World J. Surg. 28, 247–253 (2004).

    PubMed  Google Scholar 

  19. Postovit, L. M., Adams, M. A., Lash, G. E., Heaton, J. P. & Graham, C. H. Oxygen-mediated regulation of tumor cell invasiveness. Involvement of a nitric oxide signaling pathway. J. Biol. Chem. 277, 35730–35737 (2002).

    CAS  PubMed  Google Scholar 

  20. He, X. et al. Hypoxia increases heparanase-dependent tumor cell invasion, which can be inhibited by antiheparanase antibodies. Cancer Res. 64, 3928–3933 (2004).

    CAS  PubMed  Google Scholar 

  21. Buchler, P. et al. Hypoxia-inducible factor 1 regulates vascular endothelial growth factor expression in human pancreatic cancer. Pancreas 26, 56–64 (2003).

    CAS  PubMed  Google Scholar 

  22. Postovit, L. M., Adams, M. A., Lash, G. E., Heaton, J. P. & Graham, C. H. Nitric oxide-mediated regulation of hypoxia-induced B16F10 melanoma metastasis. Int. J. Cancer 108, 47–53 (2004).

    CAS  PubMed  Google Scholar 

  23. Krtolica, A. & Ludlow, J. W. Hypoxia arrests ovarian carcinoma cell cycle progression, but invasion is unaffected. Cancer Res. 56, 1168–1173 (1996).

    CAS  PubMed  Google Scholar 

  24. Schornack, P. A. & Gillies, R. J. Contributions of cell metabolism and H+ diffusion to the acidic pH of tumors. Neoplasia (New York) 5, 135–145 (2003). Determined proton production rates in breast cancer lines with low and high metastatic capability, and related these to glycolytic rate. These rates were used in a reaction–diffusion model to predict steady-state tumour pH values.

    CAS  Google Scholar 

  25. Griffiths, J. R., McIntyre, D. J., Howe, F. A. & Stubbs, M. Why are cancers acidic? A carrier-mediated diffusion model for H+ transport in the interstitial fluid. Novartis Found. Symp. 240, 46–62 (2001).

    CAS  PubMed  Google Scholar 

  26. Bhujwalla, Z. M. et al. Combined vascular and extracellular pH imaging of solid tumors. NMR Biomed. 15, 114–119 (2002). Used spectroscopic imaging to measure the spatial variations in tumour pH, and these were related to vascular perfusion measures in the same tumours.

    CAS  PubMed  Google Scholar 

  27. Krogh, A. The number and distribution of capillaries in muscles with calculations of the oxygen pressure head necessary for supplying the tissue. J. Physiol. 52, 409–415 (1919). Demonstrates the annulus of tissues that can be oxygenated by a single capillary.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Thomlinson, R. H. & Gray, L. H. The histological structure of some human lung cancers and the possible implications for radiotherapy. Br. J. Cancer 9, 539–549 (1955). Documents that necrosis in tumours occurs at distances from blood vessels and that this was consistent with the oxygen diffusion distances.

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Dewhirst, M. W., Secomb, T. W., Ong, E. T., Hsu, R. & Gross, J. F. Determination of local oxygen consumption rates in tumors. Cancer Res. 54, 3333–3336 (1994).

    CAS  PubMed  Google Scholar 

  30. Helmlinger, G., Yuan, F., Dellian, M. & Jain, R. K. Interstitial pH and pO2 gradients in solid tumors in vivo: high-resolution measurements reveal a lack of correlation. Nature Med. 3, 177–182 (1997). Despite its title, this very well conducted study documents the correlation between pH and oxygenation as they decrease with distances from feeding capillaries.

    CAS  PubMed  Google Scholar 

  31. Graeber, T. G. et al. Hypoxia-mediated selection of cells with diminished apoptotic potential in solid tumours. Nature 379, 88–91 (1996). Documents the somatic evolutionary pressure mediated by hypoxia.

    CAS  Google Scholar 

  32. Gatenby, R. A. & Gawlinski, E. T. A reaction-diffusion model of cancer invasion. Cancer Res. 56, 5745–5753 (1996). Mathematical methods and empirical evidence were used to demonstrate the acid-induced tumour-invasion model for the first time.

    CAS  PubMed  Google Scholar 

  33. Gatenby, R. A. & Vincent, T. L. An evolutionary model of carcinogenesis. Cancer Res. 63, 6212–6220 (2003). The formal mathematical development of evolutionary game theory in carcinogenesis.

    CAS  PubMed  Google Scholar 

  34. Fearon, E. R. & Vogelstein, B. A genetic model for colorectal tumorigenesis. Cell 61, 759–767 (1990). Introduced the molecular genetic changes that occur during gastrointestinal carcinogenesis and discussed the concept of clonal outgrowth in this context. There was no discussion of environmental selection pressures.

    CAS  Google Scholar 

  35. Chresand, T. J., Gillies, R. J. & Dale, B. E. Optimum fiber spacing in a hollow fiber bioreactor. Biotechnol. Bioeng. 32, 983–992 (1988).

    CAS  PubMed  Google Scholar 

  36. Secomb, T. W. et al. Theoretical simulation of oxygen transport to tumors by three-dimensional networks of microvessels. Adv. Exp. Med. Biol. 454, 629–634 (1998).

    CAS  PubMed  Google Scholar 

  37. Wykoff, C. C. et al. Expression of the hypoxia-inducible and tumor-associated carbonic anhydrases in ductal carcinoma in situ of the breast. Am. J. Pathol. 158, 1011–1019 (2001). This work shows, with histopathology, the expression of CA IX and CA XII in carcinoma in situ lesions. These two carbonic anhydrases are sensitive to hypoxia and these data are consistent with significant hypoxia in in situ lesions.

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Kimura, H. et al. Fluctuations in red cell flux in tumor microvessels can lead to transient hypoxia and reoxygenation in tumor parenchyma. Cancer Res. 56, 5522–5528 (1996).

    CAS  PubMed  Google Scholar 

  39. Hill, R. P., De Jaeger, K., Jang, A. & Cairns, R. pH, hypoxia and metastasis. Novartis Found. Symp. 240, 154–165 (2001).

    CAS  PubMed  Google Scholar 

  40. Gilead, A. & Neeman, M. Dynamic remodeling of the vascular bed precedes tumor growth: MLS ovarian carcinoma spheroids implanted in nude mice. Neoplasia (New York) 1, 226–230 (1999).

    CAS  Google Scholar 

  41. Baudelet, C. et al. Physiological noise in murine solid tumors using T2*-weighted gradient echo imaging: a marker for tumor acute hypoxia? Phys. Med. Biol. 49, 3389–3411 (2004).

    PubMed  Google Scholar 

  42. Braun, R. D., Lanzen, J. L. & Dewhirst, M. W. Fourier analysis of fluctuations of oxygen tension and blood flow in R3230Ac tumors and muscle in rats. Am. J. Physiol. 277, H551–H568 (1999).

    CAS  PubMed  Google Scholar 

  43. Dewhirst, M. W. et al. Microvascular studies on the origins of perfusion-limited hypoxia. Br. J. Cancer Suppl. 27, S247–S251 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Sonveaux, P. et al. Endothelin-1 is a critical mediator of myogenic tone in tumor arterioles: implications for cancer treatment. Cancer Res. 64, 3209–3214 (2004).

    CAS  PubMed  Google Scholar 

  45. Patan, S. et al. Vascular morphogenesis and remodeling in a human tumor xenograft: blood vessel formation and growth after ovariectomy and tumor implantation. Circ. Res. 89, 732–739 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Kiani, M. F., Pries, A. R., Hsu, L. L., Sarelius, I. H. & Cokelet, G. R. Fluctuations in microvascular blood flow parameters caused by hemodynamic mechanisms. Am. J. Physiol. 266, H1822–H1828 (1994). References 38–46 document the periodic nature of tumour oxygenation.

    CAS  PubMed  Google Scholar 

  47. Park, H. J., Lyons, J. C., Ohtsubo, T. & Song, C. W. Acidic environment causes apoptosis by increasing caspase activity. Br. J. Cancer 80, 1892–1897 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Williams, A. C., Collard, T. J. & Paraskeva, C. An acidic environment leads to p53 dependent induction of apoptosis in human adenoma and carcinoma cell lines: implications for clonal selection during colorectal carcinogenesis. Oncogene 18, 3199–3204 (1999).

    CAS  PubMed  Google Scholar 

  49. Shrode, L. D., Tapper, H. & Grinstein, S. Role of intracellular pH in proliferation, transformation, and apoptosis. J. Bioenerg. Biomembr. 29, 393–399 (1997). References 47–49 deal with pH-induced apoptosis. Grinstein's review concludes that cytoplasmic acidification is unlikely to be part of the apoptosis paradigm, but that externally lowered pH might promote apoptotic cell death.

    CAS  PubMed  Google Scholar 

  50. Lee, A. H. & Tannock, I. F. Heterogeneity of intracellular pH and of mechanisms that regulate intracellular pH in populations of cultured cells. Cancer Res. 58, 1901–1908 (1998).

    CAS  PubMed  Google Scholar 

  51. Ober, S. S. & Pardee, A. B. Intracellular pH is increased after transformation of Chinese hamster embryo fibroblasts. Proc. Natl Acad. Sci. USA 84, 2766–2770 (1987).

    CAS  PubMed  Google Scholar 

  52. McLean, L. A., Roscoe, J., Jorgensen, N. K., Gorin, F. A. & Cala, P. M. Malignant gliomas display altered pH regulation by NHE1 compared with nontransformed astrocytes. Am. J. Physiol. 278, C676–C688 (2000).

    CAS  Google Scholar 

  53. Martinez-Zaguilan, R., Lynch, R. M., Martinez, G. M. & Gillies, R. J. Vacuolar type proton ATPases are functionally expressed in the plasma membranes of human tumor cells. Am. J. Physiol. 265, c1015–c1029 (1993). References 50–53 describe mechanisms of pH regulation that are documented to be upregulated in cancers.

    CAS  PubMed  Google Scholar 

  54. Gottlieb, R. A., Giesing, H. A., Zhu, J. Y., Engler, R. L. & Babior, B. M. Cell acidification in apoptosis: granulocyte colony-stimulating factor delays programmed cell death in neutrophils by up-regulating the vacuolar H+-ATPase. Proc. Natl Acad. Sci. USA 92, 5965–5968 (1995). Demonstrates that vacuolar H+-ATPase activity is anti-apoptotic.

    CAS  PubMed  Google Scholar 

  55. Younes, M., Ertan, A., Lechago, L. V., Somoano, J. & Lechago, J. Human erythrocyte glucose transporter (Glut1) is immunohistochemically detected as a late event during malignant progression in Barrett's metaplasia. Cancer Epidemiol. Biomarkers Prev. 6, 303–305 (1997).

    CAS  PubMed  Google Scholar 

  56. Sakashita, M. et al. Glut1 expression in T1 and T2 stage colorectal carcinomas: its relationship to clinicopathological features. Eur. J. Cancer 37, 204–209 (2001).

    CAS  PubMed  Google Scholar 

  57. Grover-McKay, M., Walsh, S. A., Seftor, E. A., Thomas, P. A. & Hendrix, M. J. Role for glucose transporter 1 protein in human breast cancer. Pathol. Oncol. Res. 4, 115–120 (1998).

    CAS  PubMed  Google Scholar 

  58. Semenza, G. L. Hypoxia-inducible factor 1: master regulator of O2 homeostasis. Curr. Opin. Genet. Dev. 8, 588–594 (1998).

    CAS  PubMed  Google Scholar 

  59. Yasuda, S. et al. Hexokinase II and VEGF expression in liver tumors: correlation with hypoxia-inducible factor 1α and its significance. J. Hepatol. 40, 117–123 (2004).

    CAS  PubMed  Google Scholar 

  60. Carmeliet, P. et al. Role of HIF-1α in hypoxia-mediated apoptosis, cell proliferation and tumour angiogenesis. Nature 394, 485–490 (1998).

    CAS  Google Scholar 

  61. Robey, I., Lien, A., Welsh, S., Baggett, B. & Gillies, R. J. HIF-1α and the glycolytic phenotype in tumors. Neoplasia (in the press).

  62. Lu, H., Forbes, R. A. & Verma, A. Hypoxia-inducible factor 1 activation by aerobic glycolysis implicates the Warburg effect in carcinogenesis. J. Biol. Chem. 277, 23111–23115 (2002).

    CAS  PubMed  Google Scholar 

  63. Semenza, G. Targeting HIF-1 for cancer therapy. Nature Rev. Cancer 3, 1–13 (2003). References 58–63 describe the role of HIF1α in regulating aerobic and anaerobic glycolysis.

    Google Scholar 

  64. Semenza, G. Signal transduction to hypoxia-inducible factor 1. Biochem. Pharmacol. 64, 993–998 (2002).

    CAS  PubMed  Google Scholar 

  65. Welsh, S. J., Bellamy, W. T., Briehl, M. M. & Powis, G. The redox protein thioredoxin-1 (Trx-1) increases hypoxia-inducible factor 1α protein expression: Trx-1 overexpression results in increased vascular endothelial growth factor production and enhanced tumor angiogenesis. Cancer Res. 62, 5089–5095 (2002).

    CAS  PubMed  Google Scholar 

  66. Moeller, B. J., Cao, Y., Li, C. Y. & Dewhirst, M. W. Radiation activates HIF-1 to regulate vascular radiosensitivity in tumors: role of reoxygenation, free radicals, and stress granules. Cancer Cell 5, 429–441 (2004).

    CAS  PubMed  Google Scholar 

  67. Seagroves, T. et al. Transcription Factor HIF-1 is a necessary mediator of the Pasteur effect in mammalian cells. Mol. Cell. Biol. 21, 3436–3444 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Brand, K. Aerobic glycolysis by proliferating cells: protection against oxidative stress at the expense of energy yield. J. Bioenerg. Biomembr. 29, 355–364 (1997).

    CAS  PubMed  Google Scholar 

  69. Osthus, R. C. et al. Deregulation of glucose transporter 1 and glycolytic gene expression by c-Myc. J. Biol. Chem. 275, 21797–21800 (2000).

    CAS  PubMed  Google Scholar 

  70. Noguchi, Y. et al. Expression of facilitative glucose transporter 1 mRNA in colon cancer was not regulated by k-ras. Cancer Letters 154, 137–142 (2000).

    CAS  PubMed  Google Scholar 

  71. Chen, C., Pore, N., Behrooz, A., Ismail-Beigi, F. & Maity, A. Regulation of glut1 mRNA by hypoxia-inducible factor-1. Interaction between H-ras and hypoxia. J. Biol. Chem. 276, 9519–9525 (2001).

    CAS  Google Scholar 

  72. Mathupala, S. P., Heese, C. & Pedersen, P. L. Glucose catabolism in cancer cells. The type II hexokinase promoter contains functionally active response elements for the tumor suppressor p53. J. Biol. Chem. 272, 22776–22780 (1997).

    CAS  PubMed  Google Scholar 

  73. Goel, A., Mathupala, S. P. & Pedersen, P. L. Glucose metabolism in cancer. Evidence that demethylation events play a role in activating type II hexokinase gene expression. J. Biol. Chem. 278, 15333–15340 (2003).

    CAS  PubMed  Google Scholar 

  74. Gillies, R. J., Martinez-Zaguilan, R., Martinez, G. M., Serrano, R. & Perona, R. Tumorigenic 3T3 cells maintain an alkaline intracellular pH under physiological conditions. Proc. Natl Acad. Sci. USA 87, 7414–7418 (1990).

    CAS  PubMed  Google Scholar 

  75. Reshkin, S. J. et al. Na/H exchanger-dependent intracellular alkalinization is an early event in malignant transformation and play an essential role in the development of subsequent transformation-associated phenotypes. FASEB J. 14, 2185–2197 (2000).

    CAS  PubMed  Google Scholar 

  76. Li, X. et al. Relationship of MR-derived lactate, mobile lipids and relative blood volume for in vivo gliomas. Am. J. Neuroradiol. (in the press). Describes the observation of increased lactate in non-enhancing grade III gliomas, indicating that metabolic upregulation might precede angiogenesis.

  77. Nelson, S. J. Multivoxel magnetic resonance spectroscopy of brain tumors. Mol. Cancer Ther. 2, 497–507 (2003).

    CAS  PubMed  Google Scholar 

  78. Dafni, H., Landstrom, L., Schechter, B., Kohen, F. & Neeman, M. MRI and fluorescence microscopy of the acute vascular response to VEGF165: vasodilation, hyper-permeability and lymphatic uptake, followed by rapid inactivation of the growth factor. NMR Biomed. 15, 120–131 (2002).

    CAS  PubMed  Google Scholar 

  79. Hanahan, D. & Folkman, J. Patterns and emerging mechanisms of the angiogenic switch during tumorigenesis. Cell 86, 353–364 (1996).

    CAS  Google Scholar 

  80. Raghunand, N., Gatenby, R. A. & Gillies, R. J. Microenvironmental and cellular consequences of altered blood flow in tumors. Br. J. Radiol. 77, S11–S22 (2004).

    Google Scholar 

  81. Gillies, R. J., Raghunand, N., Karczmar, G. & Bhujwalla, Z. MR Imaging of the tumor microenvironment. J. Magn. Reson. Imaging 16, 430–450 (2002). A comprehensive review describing MRI of clinical and experimental tumours.

    PubMed  Google Scholar 

  82. Morita, T., Nagaki, T., Fukuda, I. & Okumura, K. Clastogenicity of low pH to various cultured mammalian cells. Mutat. Res. 268, 297–305 (1992).

    CAS  PubMed  Google Scholar 

  83. Ruch, R. J., Klaunig, J. E., Kerckaert, G. A. & LeBoeuf, R. A. Modification of gap junctional intercellular communication by changes in extracellular pH in syrian hamster embryo cells. Carcinogenesis 11, 909–913 (1990).

    CAS  PubMed  Google Scholar 

  84. Martinez-Zaguilan, R. et al. Acidic pH enhances the invasive behavior of human melanoma cells. Clin. Exp. Metastasis 14, 176–186 (1996).

    CAS  PubMed  Google Scholar 

  85. Schlappack, O. K., Zimmermann, A. & Hill, R. P. Glucose starvation and acidosis: effect on experimental metastasic potential, DNA content and MTX resistance of murine tumour cells. Br. J. Cancer 64, 663–670 (1991).

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Rozhin, J., Sameni, M., Ziegler, G. & Sloane, B. F. Pericellular pH affects distribution and secretion of cathepsin B in malignant cells. Cancer Res. 54, 6517–6525 (1994).

    CAS  PubMed  Google Scholar 

  87. Montcourrier, P., Silver, I., Farnoud, R., Bird, I. & Rochefort, H. Breast cancer cells have a high capacity to acidify extracellular milieu by a dual mechanism. Clin. Exp. Metastasis 15, 382–392 (1997).

    CAS  PubMed  Google Scholar 

  88. Brizel, D. M. et al. Elevated tumor lactate concentrations predict for an increased risk of metastases in head-and-neck cancer. Int. J. Radiat. Oncol. Biol. Phys. 51, 349–353 (2001).

    CAS  PubMed  Google Scholar 

  89. Walenta, S. et al. High lactate levels predict likelihood of metastases, tumor recurrence, and restricted patient survival in human cervical cancers. Cancer Res. 60, 916–921 (2000).

    CAS  PubMed  Google Scholar 

  90. Ito, S. et al. Coexpression of glucose transporter 1 and matrix metalloproteinase-2 in human cancers. J. Natl Cancer Instit. 94, 1080–1091 (2002).

    CAS  Google Scholar 

  91. Al Mehdi, A. B. et al. Intravascular origin of metastasis from the proliferation of endothelium-attached tumor cells: a new model for metastasis. Nature Med. 6, 100–102 (2000). A watershed paper describing the microenvironmental behaviour of lung metastases using a novel microscopy method. This paper challenges the paradigm that extravasation is a necessary component of the metastasis programme.

    CAS  PubMed  Google Scholar 

  92. Wong, C. W. et al. Intravascular location of breast cancer cells after spontaneous metastasis to the lung. Am. J. Pathol. 161, 749–753 (2002).

    PubMed  PubMed Central  Google Scholar 

  93. Rofstad, E. K. & Danielsen, T. Hypoxia-induced metastasis of human melanoma cells: involvement of vascular endothelial growth factor-mediated angiogenesis. Br. J. Cancer 80, 1697–1707 (1999). Provides clear evidence that pretreatment with acute hypoxia can increase the efficiency of metastasis.

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Younes, M., Lechago, L. V. & Lechago, J. Overexpression of the human erythrocyte glucose transporter occurs as a late event in human colorectal carcinogenesis and is associated with an increased incidence of lymph node metastases. Clin. Cancer Res. 2, 1151–1154 (1996).

    CAS  PubMed  Google Scholar 

  95. Yasuda, S. et al. 18F-FDG PET detection of colonic adenomas. J. Nucl. Med. 42, 989–992 (2001).

    CAS  PubMed  Google Scholar 

  96. Gatenby, R. A., Gawlinski, E. T., Tangen, C. M., Flanigan, R. C. & Crawford, E. D. The possible role of postoperative azotemia in enhanced survival of patients with metastatic renal cancer after cytoreductive nephrectomy. Cancer Res. 62, 5218–5222 (2002).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We wish to acknowledge the invaluable contributions of E. Gawlinski and T. Vincent for their efforts in the mathematical modelling that led to the insights presented here. We also thank E. Racker for stimulating this research by posing to us the question in the title.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Robert A. Gatenby.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

Related links

DATABASES

National Cancer Institute

breast cancer

cervical cancer

colon cancer

gastric cancer

head and neck cancer

oesophageal cancer

Entrez Gene

caspase-3

GLUT1

GLUT3

HRAS

HIF1α

KRAS

MYC

p53

VEGF

VHL

FURTHER INFORMATION

DMetrix digital imaging program

Glossary

HEXOKINASES

Enzymes that catalyse the transfer of phosphate from ATP to glucose to form glucose-6-phosphate. This is the first reaction in the metabolism of glucose and prevents efflux of glucose from the cell.

HYPOXIA

Refers to a low oxygen level. This means different levels to different investigators, but for radiation biologists hypoxia occurs at levels less than 0.1% oxygen in the gas phase. Normoxia refers to normal levels of oxygen (>10%) and anoxia refers to no oxygen.

WINDOW CHAMBER

A metal chamber with a glass window that is placed on the dorsal skin of an animal. This allows in vivo tumour growth to be continuously observed microscopically.

HAEMATOCRIT

A measure of the concentration of red cells in the blood. A reduced haematocrit decreases the oxygen-carrying capacity of the blood.

VASOMOTION

Rhythmic oscillations in vascular tone caused by local changes in smooth muscle.

VASCULAR REMODELLING

The active process of altering structure and arrangement in blood vessels through cell growth, cell death, cell migration and production or degradation of the extracellular matrix.

VMAX and KM

Terms from the Michaelis–Menten model. Applied to transport, Vmax is the maximum possible rate of uptake of a specific substrate. Km is the substrate concentration at which the substrate uptake is half of Vmax. Cell populations with low Km are better adapted to maintaining substrate uptake in conditions in which substrate concentrations are low.

CLASTOGENIC

Describing any substance or processes that increases alterations in the structure of chromosomes.

GAP JUNCTIONS

Linked channels through contiguous cell membranes that interconnect the cytoplasm of adjacent cells and allow direct exchange of ions and small molecules.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Gatenby, R., Gillies, R. Why do cancers have high aerobic glycolysis?. Nat Rev Cancer 4, 891–899 (2004). https://doi.org/10.1038/nrc1478

Download citation

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrc1478

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing