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:

Hypoxia and metabolism

Hypoxia, DNA repair and genetic instability

Key Points

  • The presence of intratumoural hypoxia is a negative prognostic indicator for many patients as it has been associated with increased local failure following radiotherapy and increased distant metastatic spread.

  • Hypoxia can drive the metastatic phenotype secondary to genetic instability, increased angiogenesis, decreased apoptosis and upregulation of a number of genes involved in the metastatic cascade (such as osteopontin, lysyl oxidase and vascular endothelial growth factor).

  • Both acute and chronic hypoxia exist in human tumours and these may have different biological consequences as a function of changes in hypoxia-inducible factor 1α-mediated transcription, altered protein translation and differential activation of hypoxia-associated cell cycle checkpoints.

  • Hypoxic cells can acquire a mutator phenotype that consists of decreased DNA repair, an increased mutation rate and increased chromosomal instability.

  • Defects in homologous recombination and mismatch repair have been documented in tumour cells that are exposed to chronic hypoxia.

  • Defective DNA repair in hypoxic cells could alter the sensitivity to radiotherapy and chemotherapy and render cells susceptible to molecular-targeted agents that are selectively toxic to checkpoint-deficient or repair-deficient tumour cells.

Abstract

Areas of hypoxic tumour tissue are known to be resistant to treatment and are associated with a poor clinical prognosis. There are several reasons why this might be, including the capacity of hypoxia to drive genomic instability and alter DNA damage repair pathways. Significantly, current models fail to distinguish between the complexities of the hypoxic microenvironment and the biological effects of acute hypoxia exposures versus longer-term, chronic hypoxia exposures on the transcription and translation of proteins involved in genetic stability and cell survival. Acute and chronic hypoxia might lead to different biology within the tumour and this might have a direct effect on the design of new therapies for the treatment of hypoxic tumours.

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: DNA double-strand break sensing and repair pathways.
Figure 2: DNA double-strand breaks and chromatin in irradiated hypoxic cells.
Figure 3: Hypoxia decreases RAD51 and homologous recombination.
Figure 4: Model of hypoxia-mediated genetic instability.

Similar content being viewed by others

References

  1. Blais, J. D. et al. Perk-dependent translational regulation promotes tumor cell adaptation and angiogenesis in response to hypoxic stress. Mol. Cell Biol. 26, 9517–9532 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Tatum, J. L. et al. Hypoxia: importance in tumor biology, noninvasive measurement by imaging, and value of its measurement in the management of cancer therapy. Int. J. Radiat. Biol. 82, 699–757 (2006).

    CAS  PubMed  Google Scholar 

  3. Brizel, D. M. et al. Tumor oxygenation predicts for the likelihood of distant metastases in human soft tissue sarcoma. Cancer Res. 56, 941–943 (1996).

    CAS  PubMed  Google Scholar 

  4. Chan, N., Milosevic, M. & Bristow, R. G. Tumor hypoxia, DNA repair and prostate cancer progression: new targets and new therapies. Future Oncol. 3, 329–341 (2007).

    CAS  PubMed  Google Scholar 

  5. Chaudary, N. & Hill, R. P. Hypoxia and metastasis. Clin. Cancer Res. 13, 1947–1949 (2007).

    CAS  PubMed  Google Scholar 

  6. Vaupel, P. & Mayer, A. Hypoxia in cancer: significance and impact on clinical outcome. Cancer Metastasis Rev. 26, 225–239 (2007). An excellent review of the role that hypoxia has in determining local control following therapy and metastatic spread.

    CAS  PubMed  Google Scholar 

  7. Semenza, G. L. Regulation of mammalian O2 homeostasis by hypoxia-inducible factor 1. Annu. Rev. Cell Dev. Biol. 15, 551–578 (1999).

    CAS  PubMed  Google Scholar 

  8. Salceda, S. & Caro, J. Hypoxia-inducible factor 1α (HIF-1α) protein is rapidly degraded by the ubiquitin-proteasome system under normoxic conditions. Its stabilization by hypoxia depends on redox-induced changes. J. Biol. Chem. 272, 22642–22647 (1997).

    CAS  PubMed  Google Scholar 

  9. Huang, L. E., Gu, J., Schau, M. & Bunn, H. F. Regulation of hypoxia-inducible factor 1α is mediated by an O2-dependent degradation domain via the ubiquitin-proteasome pathway. Proc. Natl Acad. Sci. USA 95, 7987–7992 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Semenza, G. L. Evaluation of HIF-1 inhibitors as anticancer agents. Drug Discov. Today 12, 853–859 (2007).

    CAS  PubMed  Google Scholar 

  11. Giaccia, A., Siim, B. G. & Johnson, R. S. HIF-1 as a target for drug development. Nature Rev. Drug Discov. 2, 803–811 (2003).

    CAS  Google Scholar 

  12. Moeller, B. J., Richardson, R. A. & Dewhirst, M. W. Hypoxia and radiotherapy: opportunities for improved outcomes in cancer treatment. Cancer Metastasis Rev. 26, 241–248 (2007).

    CAS  PubMed  Google Scholar 

  13. Harris, A. L. Hypoxia--a key regulatory factor in tumour growth. Nature Rev. Cancer 2, 38–47 (2002).

    CAS  Google Scholar 

  14. Petersen, I. Antiangiogenesis, anti-VEGF(R) and outlook. Recent Results Cancer Res. 176, 189–199 (2007).

    CAS  PubMed  Google Scholar 

  15. Chi, J. T. et al. Gene expression programs in response to hypoxia: cell type specificity and prognostic significance in human cancers. PLoS Med. 3, e47 (2006).

    PubMed  PubMed Central  Google Scholar 

  16. Koritzinsky, M. et al. The hypoxic proteome is influenced by gene-specific changes in mRNA translation. Radiother Oncol. 76, 177–186 (2005).

    CAS  PubMed  Google Scholar 

  17. Wouters, B. G. et al. Control of the hypoxic response through regulation of mRNA translation. Semin. Cell Dev. Biol. 16, 487–501 (2005).

    CAS  PubMed  Google Scholar 

  18. Koritzinsky, M. et al. Gene expression during acute and prolonged hypoxia is regulated by distinct mechanisms of translational control. EMBO J. 25, 1114–1125 (2006). Recent work delineating distinct pathways that regulate translation in response to hypoxia, thereby giving rise to differential gene expression.

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Koritzinsky, M. et al. Phosphorylation of eIF2α is required for mRNA translation inhibition and survival during moderate hypoxia. Radiother. Oncol. 83, 353–361 (2007).

    CAS  PubMed  Google Scholar 

  20. Papandreou, I. et al. Anoxia is necessary for tumor cell toxicity caused by a low-oxygen environment. Cancer Res. 65, 3171–3178 (2005).

    CAS  PubMed  Google Scholar 

  21. Bindra, R. S., Crosby, M. E. & Glazer, P. M. Regulation of DNA repair in hypoxic cancer cells. Cancer Metastasis Rev. 26, 249–260 (2007).

    CAS  PubMed  Google Scholar 

  22. Huang, L. E., Bindra, R. S., Glazer, P. M. & Harris, A. L. Hypoxia-induced genetic instability — a calculated mechanism underlying tumor progression. J. Mol. Med. 85, 139–148 (2007).

    CAS  PubMed  Google Scholar 

  23. Rofstad, E. K., Galappathi, K., Mathiesen, B. & Ruud, E. B. Fluctuating and diffusion-limited hypoxia in hypoxia-induced metastasis. Clin. Cancer Res. 13, 1971–1978 (2007).

    CAS  PubMed  Google Scholar 

  24. Hill, R. & Bristow, R. G. in The Basic Science of Oncology 4th edn (eds Tannock, I. F., Hill, R. P., Harrington, L. & Bristow, R.G.) 289–321 (McGraw-Hill, New York, 2005).

    Google Scholar 

  25. Olive, P. L. & Banath, J. P. Phosphorylation of histone H2AX as a measure of radiosensitivity. Int. J. Radiat. Oncol. Biol. Phys. 58, 331–335 (2004).

    CAS  PubMed  Google Scholar 

  26. Frankenberg-Schwager, M., Frankenberg, D. & Harbich, R. Different oxygen enhancement ratios for induced and unrejoined DNA double-strand breaks in eukaryotic cells. Radiat. Res. 128, 243–250 (1991).

    CAS  PubMed  Google Scholar 

  27. Terris, D. J. et al. Estimating DNA repair by sequential evaluation of head and neck tumor radiation sensitivity using the comet assay. Arch. Otolaryngol. Head Neck Surg. 128, 698–702 (2002).

    PubMed  Google Scholar 

  28. Wang, J., Biedermann, K. A. & Brown, J. M. Repair of DNA and chromosome breaks in cells exposed to SR 4233 under hypoxia or to ionizing radiation. Cancer Res. 52, 4473–4477 (1992).

    CAS  PubMed  Google Scholar 

  29. Brown, J. M., Evans, J. & Kovacs, M. S. The prediction of human tumor radiosensitivity in situ: an approach using chromosome aberrations detected by fluorescence in situ hybridization. Int. J. Radiat. Oncol. Biol. Phys. 24, 279–286 (1992).

    CAS  PubMed  Google Scholar 

  30. Zhang, H., Koch, C. J., Wallen, C. A. & Wheeler, K. T. Radiation-induced DNA damage in tumors and normal tissues. III. Oxygen dependence of the formation of strand breaks and DNA-protein crosslinks. Radiat. Res. 142, 163–168 (1995).

    CAS  PubMed  Google Scholar 

  31. Wardman, P. Chemical radiosensitizers for use in radiotherapy. Clin. Oncol. (R. Coll. Radiol) 19, 397–417 (2007).

    CAS  Google Scholar 

  32. Overgaard, J. Hypoxic radiosensitization: adored and ignored. J. Clin. Oncol. 25, 4066–4074 (2007). An important review of the history of the study of hypoxia in terms of predicting response to radiotherapy and radiosensitization strategies based on targeting hypoxic cells.

    PubMed  Google Scholar 

  33. Shannon, A. M., Bouchier-Hayes, D. J., Condron, C. M. & Toomey, D. Tumour hypoxia, chemotherapeutic resistance and hypoxia-related therapies. Cancer Treat. Rev. 29, 297–307 (2003).

    CAS  PubMed  Google Scholar 

  34. Minchinton, A. I. & Tannock, I. F. Drug penetration in solid tumours. Nature Rev. Cancer 6, 583–592 (2006).

    CAS  Google Scholar 

  35. Moeller, B. J. et al. Pleiotropic effects of HIF-1 blockade on tumor radiosensitivity. Cancer Cell 8, 99–110 (2005). This paper reported that fractionated radiotherapy could increase HIF1 with consequential effects on tumour cells and tumour endothelium.

    CAS  PubMed  Google Scholar 

  36. Unruh, A. et al. The hypoxia-inducible factor-1 α is a negative factor for tumor therapy. Oncogene 22, 3213–3220 (2003).

    CAS  PubMed  Google Scholar 

  37. Martinive, P. et al. Preconditioning of the tumor vasculature and tumor cells by intermittent hypoxia: implications for anticancer therapies. Cancer Res. 66, 11736–11744 (2006).

    CAS  PubMed  Google Scholar 

  38. Koritzinsky, M., Wouters, B. G., Amellem, O. & Pettersen, E. O. Cell cycle progression and radiation survival following prolonged hypoxia and re-oxygenation. Int. J. Radiat. Biol. 77, 319–328 (2001).

    CAS  PubMed  Google Scholar 

  39. Zolzer, F. & Streffer, C. Increased radiosensitivity with chronic hypoxia in four human tumor cell lines. Int. J. Radiat. Oncol. Biol. Phys. 54, 910–920 (2002).

    PubMed  Google Scholar 

  40. Shrieve, D. C. & Harris, J. W. The in vitro sensitivity of chronically hypoxic EMT6/SF cells to X-radiation and hypoxic cell radiosensitizers. Int. J. Radiat. Biol. Relat Stud Phys. Chem. Med. 48, 127–138 (1985). One of the first reports suggesting that cells that are exposed to chronic hypoxia may have a different OER to cells that are exposed to acute hypoxia

    CAS  PubMed  Google Scholar 

  41. Hockel, M. et al. Association between tumor hypoxia and malignant progression in advanced cancer of the uterine cervix. Cancer Res. 56, 4509–4515 (1996).

    CAS  PubMed  Google Scholar 

  42. Young, S. D., Marshall, R. S. & Hill, R. P. Hypoxia induces DNA overreplication and enhances metastatic potential of murine tumor cells. Proc. Natl Acad. Sci. USA 85, 9533–9537 (1988). One of the first important papers to link genetic instability, chemosensitivity and mestastasis to hypoxia

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Young, S. D. & Hill, R. P. Effects of reoxygenation on cells from hypoxic regions of solid tumors: anticancer drug sensitivity and metastatic potential. J. Natl Cancer Inst. 82, 371–380 (1990). Another of the first important papers to link genetic instability, chemosensitivity and mestastasis to hypoxia

    CAS  PubMed  Google Scholar 

  44. Cairns, R. A. & Hill, R. P. Acute hypoxia enhances spontaneous lymph node metastasis in an orthotopic murine model of human cervical carcinoma. Cancer Res. 64, 2054–2061 (2004).

    CAS  PubMed  Google Scholar 

  45. Papp-Szabo, E., Josephy, P. D. & Coomber, B. L. Microenvironmental influences on mutagenesis in mammary epithelial cells. Int. J. Cancer 116, 679–685 (2005).

    CAS  PubMed  Google Scholar 

  46. Luk, C. K., Veinot-Drebot, L., Tjan, E. & Tannock, I. F. Effect of transient hypoxia on sensitivity to doxorubicin in human and murine cell lines. J. Natl Cancer Inst. 82, 684–692 (1990).

    CAS  PubMed  Google Scholar 

  47. Rofstad, E. K. Microenvironment-induced cancer metastasis. Int. J. Radiat. Biol. 76, 589–605 (2000).

    CAS  PubMed  Google Scholar 

  48. Erler, J. T. et al. Lysyl oxidase is essential for hypoxia-induced metastasis. Nature 440, 1222–1226 (2006).

    CAS  PubMed  Google Scholar 

  49. Chan, D. A. & Giaccia, A. J. Hypoxia, gene expression, and metastasis. Cancer Metastasis Rev. 26, 333–339 (2007).

    CAS  PubMed  Google Scholar 

  50. Kaplan, R. N., Rafii, S. & Lyden, D. Preparing the “soil”: the premetastatic niche. Cancer Res. 66, 11089–11093 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Graeber, T. G. et al. Hypoxia-mediated selection of cells with diminished apoptotic potential in solid tumours. Nature 379, 88–91 (1996).

    CAS  PubMed  Google Scholar 

  52. Zhang, L. & Hill, R. P. Hypoxia enhances metastatic efficiency by up-regulating Mdm2 in KHT cells and increasing resistance to apoptosis. Cancer Res. 64, 4180–4189 (2004).

    CAS  PubMed  Google Scholar 

  53. Zhang, L. & Hill, R. P. Hypoxia enhances metastatic efficiency in HT1080 fibrosarcoma cells by increasing cell survival in lungs, not cell adhesion and invasion. Cancer Res. 67, 7789–7797 (2007).

    CAS  PubMed  Google Scholar 

  54. Graeber, T. G. et al. Hypoxia induces accumulation of p53 protein, but activation of a G1-phase checkpoint by low-oxygen conditions is independent of p53 status. Mol. Cell Biol. 14, 6264–6277 (1994).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Nordsmark, M. et al. Hypoxia in human soft tissue sarcomas: adverse impact on survival and no association with p53 mutations. Br. J. Cancer 84, 1070–1075 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Cairns, R. A., Kalliomaki, T. & Hill, R. P. Acute (cyclic) hypoxia enhances spontaneous metastasis of KHT murine tumors. Cancer Res. 61, 8903–8908 (2001).

    CAS  PubMed  Google Scholar 

  57. Kim, J. W., Gao, P. & Dang, C. V. Effects of hypoxia on tumor metabolism. Cancer Metastasis Rev. 26, 291–298 (2007).

    PubMed  Google Scholar 

  58. Bell, E. L. & Chandel, N. S. Mitochondrial oxygen sensing: regulation of hypoxia-inducible factor by mitochondrial generated reactive oxygen species. Essays Biochem. 43, 17–27 (2007).

    CAS  PubMed  Google Scholar 

  59. Guzy, R. D. & Schumacker, P. T. Oxygen sensing by mitochondria at complex III: the paradox of increased reactive oxygen species during hypoxia. Exp. Physiol. 91, 807–819 (2006).

    CAS  PubMed  Google Scholar 

  60. Gibson, S. L., Bindra, R. S. & Glazer, P. M. Hypoxia-induced phosphorylation of Chk2 in an ataxia telangiectasia mutated-dependent manner. Cancer Res. 65, 10734–10741 (2005).

    CAS  PubMed  Google Scholar 

  61. Freiberg, R. A., Hammond, E. M., Dorie, M. J., Welford, S. M. & Giaccia, A. J. DNA damage during reoxygenation elicits a Chk2-dependent checkpoint response. Mol. Cell Biol. 26, 1598–1609 (2006). References 60 and 61 showed that low O 2 leads to an ATM-dependent checkpoint and that reoxygenation leads to a CHK2-dependent checkpoint.

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Hammond, E. M., Dorie, M. J. & Giaccia, A. J. ATR/ATM targets are phosphorylated by ATR in response to hypoxia and ATM in response to reoxygenation. J. Biol. Chem. 278, 12207–12213 (2003).

    CAS  PubMed  Google Scholar 

  63. Hammer, S., To, K. K., Yoo, Y. G., Koshiji, M. & Huang, L. E. Hypoxic suppression of the cell cycle gene CDC25A in tumor cells. Cell Cycle 6, 1919–1926 (2007).

    CAS  Google Scholar 

  64. Reynolds, T. Y., Rockwell, S. & Glazer, P. M. Genetic instability induced by the tumor microenvironment. Cancer Res. 56, 5754–5757 (1996). A first paper supporting the concept that hypoxic cells have increased mutation rates.

    CAS  PubMed  Google Scholar 

  65. Li, C. Y. et al. Persistent genetic instability in cancer cells induced by non-DNA-damaging stress exposures. Cancer Res. 61, 428–432 (2001).

    CAS  PubMed  Google Scholar 

  66. Sandhu, J. K., Haqqani, A. S. & Birnboim, H. C. Effect of dietary vitamin E on spontaneous or nitric oxide donor-induced mutations in a mouse tumor model. J. Natl Cancer Inst. 92, 1429–1433 (2000).

    CAS  PubMed  Google Scholar 

  67. Banath, J. P., Sinnott, L., Larrivee, B., MacPhail, S. H. & Olive, P. L. Growth of V79 cells as xenograft tumors promotes multicellular resistance but does not increase spontaneous or radiation-induced mutant frequency. Radiat. Res. 164, 733–744 (2005).

    CAS  PubMed  Google Scholar 

  68. Weterings, E. & van Gent, D. C. The mechanism of non-homologous end-joining: a synopsis of synapsis. DNA Repair (Amst.) 3, 1425–1435 (2004).

    CAS  Google Scholar 

  69. van Gent, D. C., Hoeijmakers, J. H. & Kanaar, R. Chromosomal stability and the DNA double-stranded break connection. Nature Rev. Genet. 2, 196–206 (2001).

    CAS  PubMed  Google Scholar 

  70. Helleday, T., Lo, J., van Gent, D. C. & Engelward, B. P. DNA double-strand break repair: from mechanistic understanding to cancer treatment. DNA Repair (Amst.) 6, 923–935 (2007).

    CAS  Google Scholar 

  71. Collis, S. J., DeWeese, T. L., Jeggo, P. A. & Parker, A. R. The life and death of DNA-PK. Oncogene 24, 949–961 (2005).

    CAS  PubMed  Google Scholar 

  72. Raderschall, E. et al. Elevated levels of Rad51 recombination protein in tumor cells. Cancer Res. 62, 219–225 (2002).

    CAS  PubMed  Google Scholar 

  73. Bertrand, P., Lambert, S., Joubert, C. & Lopez, B. S. Overexpression of mammalian Rad51 does not stimulate tumorigenesis while a dominant-negative Rad51 affects centrosome fragmentation, ploidy and stimulates tumorigenesis, in p53-defective CHO cells. Oncogene 22, 7587–7592 (2003).

    CAS  PubMed  Google Scholar 

  74. Stark, J. M., Pierce, A. J., Oh, J., Pastink, A. & Jasin, M. Genetic steps of mammalian homologous repair with distinct mutagenic consequences. Mol. Cell Biol. 24, 9305–9316 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Ouyang, H. et al. Ku70 is required for DNA repair but not for T cell antigen receptor gene recombination in vivo. J. Exp. Med. 186, 921–929 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  76. Powell, S. N. & Kachnic, L. A. Roles of BRCA1 and BRCA2 in homologous recombination, DNA replication fidelity and the cellular response to ionizing radiation. Oncogene 22, 5784–5791 (2003).

    CAS  PubMed  Google Scholar 

  77. Chan, N. M. A. et al. Chronic hypoxia decreases synthesis of homologous recombination proteins to offset chemoresistance and radioresistance. Cancer Res. 68, 605–614 (2008). An important paper that shows differential HR and sensitivity to radiotherapy and chemotherapy in cells exposed to acute versus chronic hypoxia

    CAS  PubMed  Google Scholar 

  78. Bindra, R. S. & Glazer, P. M. Repression of RAD51 gene expression by E2F4/p130 complexes in hypoxia. Oncogene 26, 2048–2057 (2007).

    CAS  PubMed  Google Scholar 

  79. Sprong, D., Janssen, H. L., Vens, C. & Begg, A. C. Resistance of hypoxic cells to ionizing radiation is influenced by homologous recombination status. Int. J. Radiat. Oncol. Biol. Phys. 64, 562–572 (2006).

    CAS  PubMed  Google Scholar 

  80. Murray, D., Macann, A., Hanson, J. & Rosenberg, E. ERCC1/ERCC4 5′-endonuclease activity as a determinant of hypoxic cell radiosensitivity. Int. J. Radiat. Biol. 69, 319–327 (1996).

    CAS  PubMed  Google Scholar 

  81. Murray, D., Vallee-Lucic, L., Rosenberg, E. & Andersson, B. Sensitivity of nucleotide excision repair-deficient human cells to ionizing radiation and cyclophosphamide. Anticancer Res. 22, 21–26 (2002).

    CAS  PubMed  Google Scholar 

  82. Bindra, R. S. et al. Down-regulation of Rad51 and decreased homologous recombination in hypoxic cancer cells. Mol. Cell Biol. 24, 8504–8518 (2004). The first paper to document decreased RAD51 expression as a function of hypoxia in tissue culture and xenograft models.

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Meng, A. X. et al. Hypoxia down-regulates DNA double strand break repair gene expression in prostate cancer cells. Radiother Oncol. 76, 168–176 (2005).

    CAS  PubMed  Google Scholar 

  84. Bindra, R. S. & Glazer, P. M. Genetic instability and the tumor microenvironment: towards the concept of microenvironment-induced mutagenesis. Mutat. Res. 569, 75–85 (2005).

    CAS  PubMed  Google Scholar 

  85. Um, J. H. et al. Association of DNA-dependent protein kinase with hypoxia inducible factor-1 and its implication in resistance to anticancer drugs in hypoxic tumor cells. Exp. Mol. Med. 36, 233–242 (2004).

    CAS  PubMed  Google Scholar 

  86. To, K. K., Sedelnikova, O. A., Samons, M., Bonner, W. M. & Huang, L. E. The phosphorylation status of PAS-B distinguishes HIF-1α from HIF-2α in NBS1 repression. EMBO J. 25, 4784–4794 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  87. Mihaylova, V. T. et al. Decreased expression of the DNA mismatch repair gene Mlh1 under hypoxic stress in mammalian cells. Mol. Cell Biol. 23, 3265–3273 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Koshiji, M. et al. HIF-1α induces genetic instability by transcriptionally downregulating MutSα expression. Mol. Cell 17, 793–803 (2005). References 87 and 88 were the first papers to document decreased MMR protein expression in hypoxic cells.

    CAS  PubMed  Google Scholar 

  89. Bindra, R. S. & Glazer, P. M. Co-repression of mismatch repair gene expression by hypoxia in cancer cells: role of the Myc/Max network. Cancer Lett. 252, 93–103 (2007).

    CAS  PubMed  Google Scholar 

  90. Gordan, J. D., Bertout, J. A., Hu, C. J., Diehl, J. A. & Simon, M. C. HIF-2α promotes hypoxic cell proliferation by enhancing c-myc transcriptional activity. Cancer Cell 11, 335–347 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Huang, L. E. Carrot and stick: HIF-α engages c-Myc in hypoxic adaptation. Cell Death Differ. 11 Jan 2008 (doi: 10.1038/sj.cdd.4402302).

    CAS  Google Scholar 

  92. Liu, L. et al. Hypoxia-induced energy stress regulates mRNA translation and cell growth. Mol. Cell 21, 521–531 (2006).

    PubMed  PubMed Central  Google Scholar 

  93. Zhu, Y., McAvoy, S., Kuhn, R. & Smith, D. I. RORA, a large common fragile site gene, is involved in cellular stress response. Oncogene 25, 2901–2908 (2006).

    CAS  PubMed  Google Scholar 

  94. Coquelle, A., Pipiras, E., Toledo, F., Buttin, G. & Debatisse, M. Expression of fragile sites triggers intrachromosomal mammalian gene amplification and sets boundaries to early amplicons. Cell 89, 215–225 (1997).

    CAS  PubMed  Google Scholar 

  95. Coquelle, A., Toledo, F., Stern, S., Bieth, A. & Debatisse, M. A new role for hypoxia in tumor progression: induction of fragile site triggering genomic rearrangements and formation of complex DMs and HSRs. Mol. Cell 2, 259–265 (1998). References 94 and 95 are important papers linking hypoxia to genetic instability through increased chromosomal rearrangement, gene amplification and induction of intrachromosomal fragile sites.

    CAS  PubMed  Google Scholar 

  96. Yatabe, N. et al. HIF-1-mediated activation of telomerase in cervical cancer cells. Oncogene 23, 3708–3715 (2004).

    CAS  PubMed  Google Scholar 

  97. Anderson, C. J., Hoare, S. F., Ashcroft, M., Bilsland, A. E. & Keith, W. N. Hypoxic regulation of telomerase gene expression by transcriptional and post-transcriptional mechanisms. Oncogene 25, 61–69 (2006).

    CAS  PubMed  Google Scholar 

  98. Pandey, R., Heeger, S. & Lehner, C. F. Rapid effects of acute anoxia on spindle kinetochore interactions activate the mitotic spindle checkpoint. J. Cell Sci. 120, 2807–2818 (2007).

    CAS  PubMed  Google Scholar 

  99. Johnson, A. B. & Barton, M. C. Hypoxia-induced and stress-specific changes in chromatin structure and function. Mutat. Res. 618, 149–162 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Klein, E. A., Casey, G. & Silverman, R. Genetic susceptibility and oxidative stress in prostate cancer: integrated model with implications for prevention. Urology 68, 1145–1151 (2006).

    PubMed  Google Scholar 

  101. Gao, P. et al. HIF-dependent antitumorigenic effect of antioxidants in vivo. Cancer Cell 12, 230–238 (2007). Provocative paper showing that antioxidants may prevent cancer progression through inhibition of HIF1-mediated angiogenesis rather than inhibition of genetic instability.

    CAS  PubMed  PubMed Central  Google Scholar 

  102. Bristow, R. & Hill, R. in The Basic Science of Oncology 4th edn (eds Tannock, I., Hill, R. P, Harrington, L. & Bristow, R.) 261–288 (McGraw-Hill, New York, 2005).

    Google Scholar 

  103. Brown, J. M. & Wilson, W. R. Exploiting tumour hypoxia in cancer treatment. Nature Rev. Cancer 4, 437–447 (2004).

    CAS  Google Scholar 

  104. Evans, J. W. et al. Homologous recombination is the principal pathway for the repair of DNA damage induced by tirapazamine in mammalian cells. Cancer Res. 68, 257–265 (2008).

    CAS  PubMed  Google Scholar 

  105. Choudhury, A., Cuddihy, A. & Bristow, R. G. Radiation and new molecular agents part I: targeting ATM–ATR checkpoints, DNA repair, and the proteasome. Semin. Radiat. Oncol. 16, 51–58 (2006).

    PubMed  Google Scholar 

  106. Soldatenkov, V. A. & Potaman, V. N. DNA-binding properties of poly(ADP-ribose) polymerase: a target for anticancer therapy. Curr. Drug Targets. 5, 357–365 (2004).

    CAS  PubMed  Google Scholar 

  107. Bryant, H. E. et al. Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature 434, 913–917 (2005).

    CAS  PubMed  Google Scholar 

  108. Farmer, H. et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 434, 917–921 (2005).

    CAS  PubMed  Google Scholar 

  109. McCabe, N. et al. Deficiency in the repair of DNA damage by homologous recombination and sensitivity to poly(ADP-ribose) polymerase inhibition. Cancer Res. 66, 8109–8115 (2006). References 107, 108 and 109 showed that synthetic lethality occurred in HR-defective cells that were treated with PARP inhibitors in tumour cell models.

    CAS  PubMed  Google Scholar 

  110. Kennedy, R. D. et al. Fanconi anemia pathway-deficient tumor cells are hypersensitive to inhibition of ataxia telangiectasia mutated. J. Clin. Invest. 117, 1440–1449 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

These studies are supported by National Cancer Institute of Canada with funds raised by the Terry Fox Run. R.G.B. is a Canadian Cancer Society Research Scientist.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Robert G. Bristow.

Supplementary information

Supplementary information S1 ((movie)

A normal diploid fibroblast nucleus with pan-53BP1 staining (red) and minimal γH2AX foci (green) under non-irradiated normoxic conditions in G1 phase. See also Fig. 2a. (MOV 30800 kb)

Supplementary information S2 ((movie)

A non-irradiated anoxic fibroblast nucleus arrested in S-phase with global γH2AX activation (green) throughout the chromatin. See also Fig. 2b. (MOV 37288 kb)

Related links

Related links

DATABASES

National Cancer Institute

breast cancer

cancer of musculoskeletal tissues

cervical cancer

head and neck cancer

prostate cancer

National Cancer Institute Drug Dictionary

18F-EF5

azathioprine

cisplatin

etoposide

flavopiridol

methotrexate

misonidazole

mitomycin C

PR-104

temozolomide

tirapazamine

UCN-01

FURTHER INFORMATION

Richard P. Hill's homepage

Robert G. Bristow's homepage

Spatio-Temporal Targeting and Amplification of Radiation Response

Tumour Microenvironment Group

Glossary

Hypoxic cellular adaptation

Cells can continue to proliferate in decreasing O2 gradients by altering the transcription and translation of genes that are involved in angiogenesis and cell invasion, cell metabolism and cell survival.

Perfusion

Tumours have differential perfusion of oxygen and nutrients based on the extent and function of the intratumoural vasculature and interstitial fluid pressure.

Oxygen enhancement ratio

(OER). The ratio of radiation dose required for the same biological effect (for example, cell survival or the number of DNA breaks) in the absence versus the presence of O2. This varies for different tumour cell lines and tumour tissues and reflects the radioresistance observed in acutely anoxic cells.

Interstitial fluid pressure

(IFP). High IFP (for example, in the range 10–20 mm Hg) is characteristic many human tumours and is thought to be partly due to abnormal and highly permeable tumour vessels and decreased lymphatic drainage.

2-nitroimidazoles

Pimonidazole, misonidazole, CCI103F and EF5 are examples of 2-nitroimidazoles that are bioreduced through a series of one-electron reactions in low O2 environments and create stably bound cellular adducts. These O2-dependent adducts can then be detected in vivo using immunohistochemistry or non-invasive imaging.

Bleomycin

Bleomycin is a cytotoxic glycopeptide antibiotic used in chemotherapy. It acts by inhibiting DNA synthesis and creating oxygen-dependent DNA breaks, leading to mitotic catastrophe and cell death.

Etoposide

Etoposide is an epipodophyllotoxin that inhibits DNA synthesis by forming a complex with topoisomerase II and DNA. This complex induces DNA double-strand breaks with resulting cell cycle-dependent toxicity (that is, S and G2 phase cells).

Alkylating agents

Alkylating agents (for example, temozolomide, chlorambucil and ifosfamide) add alkyl groups to the bases of DNA, which can lead to DNA breaks and crosslinks and interference with DNA replication and transcription, all resulting in cell death.

Carboplatin

Carboplatin is an alkylating agent that covalently binds to DNA to create intra-strand and inter-strand DNA–DNA crosslinks. These crosslinks inhibit DNA synthesis and transcription leading to cell death.

Chronic obstructive pulmonary disease

(COPD). A group of diseases, including chronic bronchitis and emphysema, that are characterized by the pathological limitation of airflow in the airway that is not fully reversible.

Positron-emission tomography

(PET). PET is a non-invasive imaging technique that can create three-dimensional mapping of hypoxia within the body by detecting the γrays given off by a positron-emitting isotope (for example, 18F).

Single photon emission computed tomography

(SPECT). SPECT is a nuclear medicine tomographic imaging technique that the detects γ rays that are given off by radiopharmaceuticals (for example, 99Tc) and can provide three-dimensional and functional imaging.

Functional computed tomography

Functional computed tomography (CT) is a non-invasive imaging technique that can measure blood flow, angiogenesis and other physiologic parameters in tissues using image analysis with fast CT scanning and intravenous contrast agents.

Blood oxygenation level-dependent magnetic resonance imaging

(BOLD MRI). BOLD MRI is a non-invasive technique for indirectly measuring tissue perfusion and oxygenation based on the paramagnetic qualities of deoxyhaemoglobin. BOLD MRI does not require exogenous contrast materials and one can obtain serial images at high spatial resolution over time.

Polarographic O2 electrode measurements

A microelectrode needle can be placed into normal or tumour tissue and can determine partial pressure of O2 (pO2) measurements as it moves forward through the tissue. This can be used to compare tumour hypoxia within a cohort of patients.

Fanconi anaemia

Fanconi anemia is an autosomal recessive disorder that leads to aplastic anaemia and increased sensitivity to DNA-damaging agents. The disorder is caused by altered activity of one of at least 13 genes that encode the FANC proteins that function within a pathway as part of the DNA damage response.

Nucelotide-excision repair

This pathway recognizes bulky distortions in the DNA that occur after ultraviolet radiation or chemotherapy. Recognition of these distortions leads to the removal of a short single-stranded DNA segment that includes the lesion, creating a single-strand gap in the DNA, which is subsequently filled by a DNA polymerase.

Polysomal fractionation analysis

This technique can determine the mRNA translation efficiency of specific genes following cellular hypoxia or other stressors.

Base-excision repair

Base-excision repair uses DNA glycosylases and AP endonucleases to remove DNA bases (caused by exogenous or endogenous ROS and alkylation) in order to prevent cytotoxicity or DNA polymerase errors. Once the damaged DNA base is removed, DNA polymerase and ligase activities regenerate the DNA strand and seal the DNA.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Bristow, R., Hill, R. Hypoxia, DNA repair and genetic instability. Nat Rev Cancer 8, 180–192 (2008). https://doi.org/10.1038/nrc2344

Download citation

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrc2344

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