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:

DNA repair dysregulation from cancer driver to therapeutic target

Key Points

  • The DNA damage response (DDR) coordinates the repair of DNA and the activation of cell cycle checkpoints to arrest the cell to allow time for repair.

  • DNA is subject to a high level of endogenous damage and the DDR is essential for the maintenance of genomic stability and survival.

  • Dysregulation of the DDR can lead to genomic instability that promotes cancer development but that is exploitable with both conventional cytotoxic therapy and DDR inhibitors. Downregulated DDR pathways render the tumour sensitive to specific cytotoxics and some DDR inhibitors. Upregulated DDR pathways confer therapeutic resistance.

  • Inhibitors of the DDR have been developed to overcome resistance and to augment the activity of conventional therapy.

  • Loss of a DDR pathway can lead to dependence on a compensatory pathway, and targeting this second pathway may render endogenous DNA damage cytotoxic by a process termed synthetic lethality, which will be tumour-specific because the normal tissues in the animal (or person) will have functional DNA repair.

  • Despite promising preclinical data combining DDR inhibitors with conventional cytotoxic agents, these combinations have been less successful in the clinic and are often associated with toxicity. Exploitation of DDR defects by synthetic lethality is a more promising approach. Clinical data on the use of poly(ADP-ribose) polymerase (PARP) inhibitors in homologous recombination repair (HRR)-defective tumours are encouraging.

  • Robust and validated biomarkers to identify DDR defects that are exploitable by both conventional cytotoxic therapy and agents targeting the DDR are needed to effectively stratify patients.

Abstract

Dysregulation of DNA damage repair and signalling to cell cycle checkpoints, known as the DNA damage response (DDR), is associated with a predisposition to cancer and affects responses to DNA-damaging anticancer therapy. Dysfunction of one DNA repair pathway may be compensated for by the function of another compensatory DDR pathway, which may be increased and contribute to resistance to DNA-damaging chemotherapy and radiotherapy. Therefore, DDR pathways make an ideal target for therapeutic intervention; first, to prevent or reverse therapy resistance; and second, using a synthetic lethal approach to specifically kill cancer cells that are dependent on a compensatory DNA repair pathway for survival in the context of cancer-associated oxidative and replicative stress. These hypotheses are currently being tested in the laboratory and are being translated into clinical studies

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: Sources of DNA damage and their repair.
Figure 2: Base excision repair.
Figure 3: Nucleotide excision repair.
Figure 4: Mismatch repair.
Figure 5: DNA double-strand break and interstrand crosslink repair.
Figure 6: Signalling DNA damage to cell cycle checkpoints.
Figure 7: Synthetic lethality.

Similar content being viewed by others

References

  1. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell 144, 646–674 (2011). An excellent review highlighting enabling characteristics, as well as emerging hallmarks, of cancer.

    Article  CAS  PubMed  Google Scholar 

  2. Bartkova, J. et al. DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature 434, 864–870 (2005).

    Article  CAS  PubMed  Google Scholar 

  3. Gorgoulis, V. G. et al. Activation of the DNA damage checkpoint and genomic instability in human precancerous lesions. Nature 434, 907–913 (2005) References 2 and 3 show that early in tumorigenesis oncogenic stress activates the DDR, which acts as a temporary barrier to progression, but defects in the DDR eventually arise allowing full-blown cancer to develop.

    Article  CAS  PubMed  Google Scholar 

  4. Tubbs, J. L., Pegg, A. E. & Tainer, J. A. DNA binding, nucleotide flipping, and the helix-turn-helix motif in base repair by O6-alkylguanine-DNA alkyltransferase and its implications for cancer chemotherapy. DNA Repair 6, 1100–1115 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Tricker, A. R. & Preussmann, R. Carcinogenic N-nitrosamines in the diet: occurrence, formation, mechanisms and carcinogenic potential. Mutat. Res. 259, 277–289 (1991).

    Article  CAS  PubMed  Google Scholar 

  6. Saffhill, R., Margison, G. P. & O'Connor, P. J. Mechanisms of carcinogenesis induced by alkylating agents. Biochim. Biophys. Acta 823, 111–145 (1985).

    CAS  PubMed  Google Scholar 

  7. Zaidi, N. H., Liu, L. & Gerson, S. L. Quantitative immunohistochemical estimates of O6-alkylguanine-DNA alkyltransferase expression in normal and malignant human colon. Clin. Cancer Res. 2, 577–584 (1996).

    CAS  PubMed  Google Scholar 

  8. Pegg, A. E. Mammalian O6-alkylguanine-DNA alkyltransferase: regulation and importance in response to alkylating carcinogenesis and therapeutic agents. Cancer Res. 50, 6119–6129 (1990).

    CAS  PubMed  Google Scholar 

  9. Rabik, C. A., Njoku, M. C. & Dolan, M. E. Inactivation of O6-alkylguanine DNA alkyltransferase as a means to enhance chemotherapy. Cancer Treat. Rev. 32, 261–276 (2006).

    Article  CAS  PubMed  Google Scholar 

  10. Friedman, H. S. et al. Phase I trial of O6-benzylguanine for patients undergoing surgery for malignant glioma. J. Clin. Oncol. 16, 3570–3575 (1998). The first clinical trial of a DNA repair inhibitor supported by pharmacodynamic measurement of the impact on O6-methylguanine levels in tumour material.

    Article  CAS  PubMed  Google Scholar 

  11. Ranson, M. et al. Lomeguatrib, a potent inhibitor of O6-alkylguanine-DNA-alkyltransferase: phase I safety, pharmacodynamic, and pharmacokinetic trial and evaluation in combination with temozolomide in patients with advanced solid tumors. Clin. Cancer Res. 12, 1577–1584 (2006).

    Article  CAS  PubMed  Google Scholar 

  12. Watson, A. J. et al. Tumor O6-methylguanine-DNA methyltransferase inactivation by oral lomeguatrib Clin. Cancer Res. 16, 743–749 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Esteller, M., Hamilton, S. R., Burger, P. C., Baylin, S. B. & Herman, J. G. Inactivation of the DNA repair gene O6-methylguanine-DNAmethyltransferase by promoter hypermethylation is a common event in primary human neoplasia. Cancer Res. 59, 793–797 (1999).

    CAS  PubMed  Google Scholar 

  14. Hegi, M. E. et al. MGMT gene silencing and benefit from temozolomide in glioblastoma. N. Engl. J. Med. 352, 997–1003 (2005). The demonstration in a study of >200 patients that patients with epigenetic silencing of MGMT had a significantly better response to TMZ and radiotherapy.

    Article  CAS  PubMed  Google Scholar 

  15. Lindahl, T. Instability and decay of the primary structure of DNA. Nature 362, 709–715 (1993). Review of the extent of endogenous and environmental damage to DNA, linking it to ageing and cancer.

    Article  CAS  PubMed  Google Scholar 

  16. van Loon, B., Markkanen, E. & Hübscher, U. Oxygen as a friend and enemy: how to combat the mutational potential of 8-oxo-guanine. DNA Repair 9, 604–616 (2010).

    Article  CAS  PubMed  Google Scholar 

  17. Wiseman, H. & Halliwell, B. Damage to DNA by reactive oxygen and nitrogen species: role in inflammatory disease and progression to cancer. Biochem. J. 313, 17–29 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. De Vos, M., Schreiber, V. & Dantzer, F. The diverse roles and clinical relevance of PARPs in DNA damage repair: current state of the art. Biochem. Pharmacol. 84, 137–146 (2012).

    Article  CAS  PubMed  Google Scholar 

  19. Schreiber, V., Dantzer, F., Ame, J. C. & de Murcia, G. Poly(ADP-ribose): novel functions for an old molecule. Nature Rev. Mol. Cell Biol. 7, 517–528 (2006).

    Article  CAS  Google Scholar 

  20. El-Khamisy, S. F., Masutani, M., Suzuki, H. & Caldecott, K. W. A requirement for PARP-1 for the assembly or stability of XRCC1 nuclear foci at sites of oxidative DNA damage. Nucleic Acids Res. 31, 5526–5533 (2003). An elegant study demonstrating that ADP-ribose polymer formation at the site of the DNA break is necessary to recruit XRCC1, the scaffold protein of BER.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Plo, I. et al. Association of XRCC1 and tyrosyl DNA phosphodiesterase Tdp 1 for the repair of topoisomerase I-mediated DNA lesions. DNA Repair 2, 1087–1100 (2003).

    Article  CAS  PubMed  Google Scholar 

  22. Larsen, E., Meza, T. J., Kleppa, L. & Klungland, A. Organ and cell specificity of base excision repair mutants in mice. Mutat. Res. 614, 56–68 (2007).

    Article  CAS  PubMed  Google Scholar 

  23. Sweasy, J. B., Lang, T. & DiMaio, D. Is base excision repair a tumour suppressor mechanism? Cell Cycle 5, 250–259 (2006).

    Article  CAS  PubMed  Google Scholar 

  24. Abbotts, R. & Madhusudan, S. Human AP endonuclease 1 (APE1): from mechanistic insights to druggable target in cancer. Cancer Treat. Rev. 36, 425–435 (2010).

    Article  CAS  PubMed  Google Scholar 

  25. Hirai, K., Ueda, K. & Hayaishi, O. Aberration of poly(adenosine diphosphate-ribose) metabolism in human colon adenomatous polyps and cancers. Cancer Res. 43, 3441–3446 (1983).

    CAS  PubMed  Google Scholar 

  26. Zaremba, T. et al. Poly(ADP-ribose) polymerase-1 polymorphisms, expression and activity in selected human tumour cell lines. Br. J. Cancer 21, 256–262 (2009).

    Article  CAS  Google Scholar 

  27. Rouleau, M., Patel, A., Hendzel, M. J., Kaufmann, S. H. & Poirier, G. G. PARP inhibition: PARP1 and beyond. Nature Rev. Cancer 10, 293–301 (2010).

    Article  CAS  Google Scholar 

  28. Taverna, P. et al. Methoxyamine potentiates DNA single strand breaks and double strand breaks induced by temozolomide in colon cancer cells. Mutat. Res. 485, 269–281 (2001). The early identification that inhibition of APE1 endonuclease could increase TMZ cytotoxicity.

    Article  CAS  PubMed  Google Scholar 

  29. Luo, M. & Kelley, M. R. Inhibition of the human apurinic/apyrimidinic endonuclease (APE1) repair activity and sensitization of breast cancer cells to DNA alkylating agents with lucanthone. Anticancer Res. 24, 2127–2134 (2004).

    CAS  PubMed  Google Scholar 

  30. Del Rowe, J. D. et al. Accelerated regression of brain metastases in patients receiving whole brain radiation and the topoisomerase II inhibitor, lucanthone. Int. J. Radiat. Oncol. Biol. Phys. 43, 89–93 (1999).

    Article  CAS  PubMed  Google Scholar 

  31. Mohammed, M. Z. et al. Development and evaluation of human AP endonuclease inhibitors in melanoma and glioma cell lines. Br. J. Cancer 104, 6536–6563 (2011).

    Article  CAS  Google Scholar 

  32. McNeill, D. R. & Wilson, D. M. A dominant-negative form of the major human abasic endonuclease enhances cellular sensitivity to laboratory and clinical DNA-damaging agents. Mol. Cancer Res. 5, 61–70 (2007).

    Article  CAS  PubMed  Google Scholar 

  33. Bowman K. J., White, A., Golding, B. T., Griffin, R. J. & Curtin, N. J. Potentiation of anticancer agent cytotoxicity by the potent poly(ADP-ribose) polymerase inhibitors, NU1025 and NU1064. Br. J. Cancer 78, 1269–1277 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Calabrese, C. R. et al. Preclinical evaluation of a novel poly(ADP-ribose) polymerase-1 (PARP-1) inhibitor, AG14361, with significant anticancer chemo- and radio-sensitization activity. J. Natl Cancer Inst. 96, 56–67 (2004). The first demonstration that a potent PARPi could enhance the antitumour activity of IR and topoisomerase I poisons in xenograft models, with complete durable regressions observed in combination with TMZ.

    Article  CAS  PubMed  Google Scholar 

  35. Plummer, R. et al. Phase I study of the poly(ADP-ribose) polymerase inhibitor, AG014699, in combination with temozolomide in patients with advanced solid tumors. Clin. Cancer Res. 14, 7917–7923 (2008). The first clinical trial of a PARPi in cancer patients.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Plummer, R. et al. First and final report of a phase II study of the poly(ADP-ribose) polymerase (PARP) inhibitor, AG014699, in combination with temozolomide (TMZ) in patients with metastatic malignant melanoma (MM). J. Clin. Oncol. Abstr. 24, 8013 (2006).

    Google Scholar 

  37. Zhang, Y. W. et al. Poly(ADP-ribose) polymerase and XPF-ERCC1 participate in distinct pathways for the repair of topoisomerase I-induced DNA damage in mammalian cells. Nucleic Acids Res. 39, 3128–3140 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Drew, Y. et al. Therapeutic potential of poly(ADP-ribose) polymerase inhibitor AG014699 in human cancers with mutated or methylated BRCA1 or BRCA2. J. Natl Cancer Inst. 103, 334–346 (2011).

    Article  CAS  PubMed  Google Scholar 

  39. Thomas, H. D. et al. Preclinical selection of a novel poly(ADP-ribose) polymerase (PARP) inhibitor for clinical trial. Mol. Cancer Ther. 6, 945–956 (2007).

    Article  CAS  PubMed  Google Scholar 

  40. Mehta, M. P. et al. Phase I safety and pharmacokinetic (PK) study of veliparib in combination with whole brain radiation therapy (WBRT) in patients (pts) with brain metastases. J. Clin. Oncol. Abstr. 30, 2013 (2012).

    Article  CAS  Google Scholar 

  41. Naegeli, H. & Sugasawa, K. The xeroderma pigmentosum pathway: decision tree analysis of DNA quality. DNA Repair 10, 673–683 (2011).

    Article  CAS  PubMed  Google Scholar 

  42. Andressoo, J. O., Hoeijmakers, J. H. & de Waard, H. Nucleotide excision repair and its connection with cancer and ageing. Adv. Exp. Med. Biol. 570, 45–83 (2005).

    Article  CAS  PubMed  Google Scholar 

  43. Koberle, B. et al. DNA repair capacity and cisplatin sensitivity of human testis tumour cells. Int. J. Cancer 70, 551–555 (1997).

    Article  CAS  PubMed  Google Scholar 

  44. Usanova, S. et al. Cisplatin sensitivity of testis tumour cells is due to deficiency in interstrand-crosslink repair and low ERCC1-XPF expression. Mol. Cancer 9, 248 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  45. Lange, S. S., Takata, K. & Wood, R. D. DNA polymerases and cancer. Nature Rev. Cancer 11, 96–110 (2011).

    Article  CAS  Google Scholar 

  46. Mizushina, Y. et al. 3-O-methylfunicone, a selective inhibitor of mammalian Y-family DNA polymerases from an Australian sea salt fungal strain. Mar. Drugs 7, 624–639 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Dorjsuren, D. et al. A real-time fluorescence method for enzymatic characterization of specialized human DNA polymerases. Nucleic Acids Res. 37, e128 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Umar, A. et al. Revised Bethesda Guidelines for hereditary nonpolyposis colorectal cancer (Lynch syndrome) and microsatellite instability. J. Natl Cancer Inst. 96, 261–268 (2004).

    Article  CAS  PubMed  Google Scholar 

  49. Wu, X., Xu, Y., Chai, W. & Her, C. Causal link between microsatellite instability and hMRE11 dysfunction in human cancers. Mol. Cancer Res. 9, 1443–1448 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Ham, M. F. et al. Impairment of double-strand breaks repair and aberrant splicing of ATM and MRE11 in leukemia-lymphoma cell lines with microsatellite instability. Cancer Sci. 97, 226–234 (2006). Demonstration of the impact of MMR defects on the function of HRR pathways.

    Article  CAS  PubMed  Google Scholar 

  51. Karran, P. & Bignami, M. DNA damage tolerance, mismatch repair and genome instability. Bioessays 16, 833–839 (1994).

    Article  CAS  PubMed  Google Scholar 

  52. Fordham, S. E., Matheson, E. C., Scott, K., Irving, J. A. & Allan, J. M. DNA mismatch repair status affects cellular response to Ara-C and other anti-leukemic nucleoside analogs. Leukemia 25, 1046–1049 (2011).

    Article  CAS  PubMed  Google Scholar 

  53. Plumb, J. A., Strathdee, G., Sludden, J., Kaye, S. B. & Brown, R. Reversal of drug resistance in human tumor xenografts by 2′-deoxy-5-azacytidine-induced demethylation of the hMLH1 gene promoter. Cancer Res. 60, 6039–6044 (2000).

    CAS  PubMed  Google Scholar 

  54. Vilenchik, M. M. & Knidson, A. G. Endogenous DNA double-strand breaks: production, fidelity of repair, induction of cancer. Proc. Natl Acad. Sci. USA 100, 12871–12876 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Shrivastav, M., De Haro, L. P. & Nickoloff, J. A. Regulation of DNA double-strand break repair pathway choice. Cell Res. 18, 134–147 (2008).

    Article  CAS  PubMed  Google Scholar 

  56. Mahaney, B. L., Meek, K. & Lees-Miller, S. P. Repair of ionizing radiation-induced DNA double-strand breaks by non-homologous end-joining. Biochem. J. 417, 639–650 (2009).

    Article  CAS  PubMed  Google Scholar 

  57. Beucher, A. et al. ATM and Artemis promote homologous recombination of radiation-induced DNA double-strand breaks in G2. EMBO J. 28, 3413–3427 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. McClendon, A. K. & Osheroff, N. DNA topoisomerase II, genotoxicity, cancer. Mutat. Res. 623, 83–97 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Rosenzweig, K. E., Youmell, M. B., Palayoor, S. T. & Price, B. D. Radiosensitization of human tumor cells by the phosphatidylinositol3-kinase inhibitors wortmannin and LY294002 correlates with inhibition of DNA-dependent protein kinase and prolonged G2-M delay. Clin. Cancer Res. 3, 1149–1156 (1997). Early demonstration that PI3K inhibitors that also inhibit purified and cellular DNA-PKcs are radiosensitizers.

    CAS  PubMed  Google Scholar 

  60. Boulton, S., Kyle, S. & Durkacz, B. W. Mechanisms of enhancement of cytotoxicity in etoposide and ionising radiation-treated cells by the protein kinase inhibitor wortmannin. Eur. J. Cancer 36, 535–541 (2000).

    Article  CAS  PubMed  Google Scholar 

  61. Hardcastle, I. R. et al. Discovery of potent chromen-4-one inhibitors of the DNA-dependent protein kinase (DNA-PK) using a small-molecule library approach. J. Med. Chem. 48, 7829–7846 (2005).

    Article  CAS  PubMed  Google Scholar 

  62. Shinohara, E. T. et al. DNA-dependent protein kinase is a molecular target for the development of noncytotoxic radiation-sensitizing drugs. Cancer Res. 65, 4987–4992 (2005).

    Article  CAS  PubMed  Google Scholar 

  63. Zhao, Y. et al. Preclinical evaluation of a potent novel DNA-dependent protein kinase inhibitor NU7441. Cancer Res. 66, 5354–5362 (2006). The first DNA-PKcs inhibitor that was demonstrated to induce chemosensitization in tumour-bearing mice.

    Article  CAS  PubMed  Google Scholar 

  64. Munck, J. M. et al. Chemosensitisation of cancer cells by KU-0060648; a dual inhibitor of DNA-PK and PI-3K. Mol. Can. Ther. 11, 1789–1798 (2012).

    Article  CAS  Google Scholar 

  65. Willmore, E. et al. A novel DNA-dependent protein kinase inhibitor, NU7026, potentiates the cytotoxicity of topoisomerase II poisons used in the treatment of leukemia. Blood 103, 4659–4665 (2004).

    Article  CAS  PubMed  Google Scholar 

  66. Feeney, K. M., Wasson, C. W. & Parish, J. L. Cohesin: a regulator of genome integrity and gene expression. Biochem. J. 428, 147–161 (2010).

    Article  CAS  PubMed  Google Scholar 

  67. Saleh-Gohari, N. et al. Spontaneous homologous recombination is induced by collapsed replication forks that are caused by endogenous DNA single-strand breaks. Mol.Cell Biol. 25, 7158–7169 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Deans, A. J. & West, S. C. DNA interstrand crosslink repair and cancer. Nature Rev. Cancer 11, 467–480 (2011).

    Article  CAS  Google Scholar 

  69. Evers, B., Helleday, T. & Jonkers, J. Targeting homologous recombination repair defects in cancer. Trends Pharmacol.Sci. 31, 372–380 (2010).

    Article  CAS  PubMed  Google Scholar 

  70. Dupré, A. et al. A forward chemical genetic screen reveals an inhibitor of the Mre11-Rad50-Nbs1 complex. Nature Chem. Biol. 4, 119–125 (2008).

    Article  CAS  Google Scholar 

  71. Rass, E. et al. Role of Mre11 in chromosomal nonhomologous end joining in mammalian cells. Nature Struct. Mol. Biol. 16, 819–824 (2009).

    Article  CAS  Google Scholar 

  72. Aloyz, R. et al. Imatinib sensitizes CLL lymphocytes to chlorambucil. Leukemia. 18, 409–414 (2004).

    Article  CAS  PubMed  Google Scholar 

  73. Choudhury, A. et al. Targeting homologous recombination using imatinib results in enhanced tumor cell chemosensitivity and radiosensitivity. Mol. Cancer Ther. 8, 203–213 (2009).

    Article  CAS  PubMed  Google Scholar 

  74. Dernheimer, F.A. & Kastan, M,B. Multiple roles of ATM in monitoring and maintaining DNA integrity. Febs Letts. 584, 3675–3681 (2010).

    Article  CAS  Google Scholar 

  75. Kastan, M. B. & Bartek, J. Cell cycle checkpoints and cancer. Nature 432, 316–323 (2004).

    Article  CAS  PubMed  Google Scholar 

  76. Bouwman, P. et al. 53BP1 loss rescues BRCA1 deficiency and is associated with triple-negative and BRCA-mutated breast cancers. Nature Struct. Mol. Biol. 17, 688–695 (2010). Novel observation that deletion of a component of NHEJ restores HRR function in BRCA1-mutant cells.

    Article  CAS  Google Scholar 

  77. Wang, H., Wang, H., Powel, S. N., Iliakis, G. & Wang, Y. ATR affecting cell radiosensitivity is dependent on homologous recombination but independent of non-homologous end joining. Cancer Res. 64, 7139–7143 (2004).

    Article  CAS  PubMed  Google Scholar 

  78. Flynn, R. L. & Zou, L. ATR: a master conductor of cellular responses to DNA replication stress. Trends Biochem. Sci. 36, 133–140 (2011).

    Article  CAS  PubMed  Google Scholar 

  79. Lewis, K. A. et al. Heterozygous ATR mutations in mismatch repair-deficient cancer cells have functional significance. Cancer Res. 65, 7091–7095 (2005).

    Article  CAS  PubMed  Google Scholar 

  80. Matsuoka, S. et al. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 316, 1160–1166 (2007). Proteomic analysis of phosphorylation networks downstream of DNA damage-activated ATM and ATR that identifies >700 proteins, which confirmed many targets of the DDR pathway and several new ones.

    Article  CAS  PubMed  Google Scholar 

  81. Hickson, I. et al. Identification and characterization of a novel and specific inhibitor of the ataxia-telangiectasia mutated kinase ATM. Cancer Res. 64, 9152–9159 (2004).

    Article  CAS  PubMed  Google Scholar 

  82. Bartkova, J. et al. Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature. 444, 633–637 (2006).

    Article  CAS  PubMed  Google Scholar 

  83. Massague, J. G1 cell cycle control and cancer. Nature 432, 298–306 (2004).

    Article  CAS  PubMed  Google Scholar 

  84. Cliby, W. A. et al. Overexpression of a kinase-inactive ATR protein causes sensitivity to DNA damaging agents and defects in cell cycle checkpoints. EMBO J. 17, 159–169 (1998). The first demonstration of the potential chemosensitization effect of ATR activation using an ATR kinase-dead dominant-negative approach.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Carrassa, L. Broggini, M., Erba, E. & Damia, G. Chk1, but not Chk2, is involved in the cellular response to DNA damaging agents: differential activity in cells expressing or not p53. Cell Cycle 3, 1177–1181 (2004).

    Article  CAS  PubMed  Google Scholar 

  86. Nghiem, P., Park, P. K., Kim, Y.-S., Vaziri, C. & Schreiber, S. L. ATR inhibition selectively sensitizes G1 checkpoint-deficient cells to lethal premature chromatin condensation. Proc. Natl Acad. Sci. USA 98, 9092–9097 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Garrett, M. D. & Collins, I. Anticancer therapy with checkpoint inhibitors: what, where and when? Trends Pharm. Sci. 32, 308–316 (2011).

    Article  CAS  PubMed  Google Scholar 

  88. Blasina, A. et al. Breaching the DNA damage checkpoint via PF-00477736, a novel small-molecule inhibitor of checkpoint kinase 1. Mol. Cancer Ther. 7, 2394–2404 (2008).

    Article  CAS  PubMed  Google Scholar 

  89. Zabludoff, S. D. et al. AZD7762, a novel checkpoint kinase inhibitor, drives checkpoint abrogation and potentiates DNA-targeted therapies. Mol. Cancer Ther. 7, 2955–2966 (2008).

    Article  CAS  PubMed  Google Scholar 

  90. Matthews, D. J. et al. Pharmacological abrogation of S-phase checkpoint enhances the anti-tumor activity of gemcitabine in vivo. Cell Cycle 6, 104–110 (2007).

    Article  CAS  PubMed  Google Scholar 

  91. Chen, T., Stephens, P. A., Middleton, F. K. & Curtin, N. J. Targeting the S and G2 checkpoint to treat cancer. Drug Discov. Today 17, 194–202 (2012).

    Article  CAS  PubMed  Google Scholar 

  92. Sarkaria, J. N. et al. Inhibition of ATM and ATR kinase activities by the radiosensitizing agent, caffeine. Cancer Res. 59, 4375–4382 (1999).

    CAS  PubMed  Google Scholar 

  93. Toledo, L. I. et al. A cell-based screen identifies ATR inhibitors with synthetic lethal properties for cancer-associated mutations. Nature Struct. Mol. Biol. 18, 721–727 (2011).

    Article  CAS  Google Scholar 

  94. Reaper, P. M. et al. Selective killing of ATM- or p53-deficient cancer cells through inhibition of ATR. Nature Chem. Biol. 7, 428–430 (2011). The first report of a potent and specific small-molecule ATR inhibitor.

    Article  CAS  Google Scholar 

  95. Peasland, A., et al. Identification and evaluation of a potent novel ATR inhibitor, NU6027, in breast and ovarian cancer cell lines. Br. J. Cancer 105, 372–381 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Hartwell, L. H., Szankasi, P., Roberts, C. J., Murray, A. W. & Friend, S. H. Integrating Genetic approaches into the discovery of anticancer drugs. Science 278, 1064–1068 (1997). The first paper to identify the potential use of synthetic lethality in cancer.

    Article  CAS  PubMed  Google Scholar 

  97. Dolma, S., Lessnick, S. L., Hahn, W. C. & Stockwell, B. R. Identification of genotype-selective antitumor agents using synthetic lethal chemical screening in engineered human tumor cells. Cancer Cell 3, 285–289 (2003).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  99. Farmer, H. et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 434, 917–921 (2005). References 98 and 99 were the first to demonstrate that PARPis are synthetically lethal with HRR-defective cells and xenografts.

    Article  CAS  PubMed  Google Scholar 

  100. Helleday, T. The underlying mechanism for the PARP & BRCA synthetic lethality: clearing up the misunderstandings. Mol. Oncol. 5, 387–393 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Chan, N. et al. Contextual synthetic lethality of cancer cell kill based on the tumor microenvironment. Cancer Res. 70, 8045–8054 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Krawczyk, P. M. et al. Mild hyperthermia inhibits homologous recombination, induces BRCA2 degradation, and sensitizes cancer cells to poly (ADP-ribose) polymerase-1 inhibition. Proc. Natl Acad. Sci. USA 108, 9851–9856 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Fong, P. C. et al. Poly(ADP)-ribose polymerase inhibition: frequent durable responses in BRCA carrier ovarian cancer correlating with platinum-free interval. J. Clin. Oncol. 28, 2512–2519 (2010). Report of the first clinical trial of a PARPi in patients with BRCA-associated tumours, demonstrating a good response rate with minimal toxicity.

    Article  CAS  PubMed  Google Scholar 

  104. Gelmon, K. A. et al. Olaparib in patients with recurrent high-grade serous or poorly differentiated ovarian carcinoma or triple-negative breast cancer: a phase 2, multicentre, open-label, non-randomised study. Lancet Oncol. 12, 852–861 (2011).

    Article  CAS  PubMed  Google Scholar 

  105. 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).

    Article  CAS  PubMed  Google Scholar 

  106. Sultana, R. et al. Synthetic lethal targeting of DNA double-strand break repair deficient cells by human apurinic/apyrimidinic endonuclease inhibitors. Int. J. Cancer 131, 2433–2444 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Sakai, W. et al. Secondary mutations as a mechanism of cisplatin resistance in Brca2-mutated cancers. Nature. 451, 1116–1120 (2008). Demonstration of resistance to PARPi and cisplatin in BRCA-mutant cells by virtue of secondary mutations restoring BRCA protein and HRR function.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Patel, A. G., Sarkaria, J. N. & Kaufmann, S. H. Nonhomologous end joining drives poly(ADP-ribose) polymerase (PARP) inhibitor lethality in homologous recombination-deficient cells. Proc. Natl Acad.Sci. USA 108, 3406–3411 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Adamo, A. et al. Preventing nonhomologous end joining suppresses DNA repair defects of Fanconi anemia. Mol. Cell. 39, 25–35 (2010).

    Article  CAS  PubMed  Google Scholar 

  110. Bartkova, J. et al. DNA damage response mediators MDC1 and 53BP1: constitutive activation and aberrant loss in breast and lung cancer, but not in testicular germ cell tumours. Oncogene 26, 7414–7422 (2007).

    Article  CAS  PubMed  Google Scholar 

  111. Weston, V. J. et al. The PARP inhibitor olaparib induces significant killing of ATM-deficient lymphoid tumor cells in vitro and in vivo. Blood 116, 4578–4587 (2010).

    Article  CAS  PubMed  Google Scholar 

  112. Williamson, C. T. et al. ATM deficiency sensitizes mantle cell lymphoma cells to poly(ADP-ribose) polymerase-1 inhibitors. Mol. Cancer Ther. 9, 347–357 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Vilar, E. et al. MRE11 deficiency increases sensitivity to poly(ADP-ribose) polymerase inhibition in microsatellite unstable colorectal cancers. Cancer Res. 71, 2632–2642 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Martin, S. A. et al. DNA polymerases as potential therapeutic targets for cancers deficient in the DNA mismatch repair proteins MSH2 or MLH1. Cancer Cell 17, 235–248 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Johnson, N. et al. Compromised CDK1 activity sensitizes BRCA-proficient cancers to PARP inhibition. Nature Med. 17, 875–883 (2011). Demonstration of the synthetic lethality of CDK1 inhibition and PARP inhibition in tumours, whilst sparing normal tissues, in an animal model.

    Article  CAS  PubMed  Google Scholar 

  116. Mitchell, C. et al. Poly(ADP-ribose) polymerase 1 modulates the lethality of CHK1 inhibitors in carcinoma cells. Mol. Pharmacol. 78, 909–917 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Höglund, A., Strömvall, K., Li, Y., Forshell, L. P. & Nilsson, J. A. Chk2 deficiency in Myc overexpressing lymphoma cells elicits a synergistic lethal response in combination with PARP inhibition. Cell Cycle. 10, 3598–3607 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  118. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Chen, C. C., Kennedy, R. D., Sidi, S., Look, A. T. & D'Andrea, A. CHK1 inhibition as a strategy for targeting Fanconi Anemia (FA) DNA repair pathway deficient tumors. Mol. Cancer 8, 24 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  120. Gilad, O. et al. Combining ATR suppression with oncogenic Ras synergistically increases genomic instability, causing synthetic lethality or tumorigenesis in a dosage-dependent manner. Cancer Res. 70, 9693–9702 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Ferrao, P. T. Bukczynska, E. P., Johnstone, R. W. & McArthur, G. A. Efficacy of CHK inhibitors as single agents in MYC-driven lymphoma cells. Oncogene 31, 1661–1672 (2011).

    Article  PubMed  CAS  Google Scholar 

  122. Brenner, J. C. et al. Mechanistic rationale for inhibition of poly(ADP-ribose) polymerase in ETS gene fusion-positive prostate cancer. Cancer Cell 19, 664–678 (2011). Demonstration that ETS fusion genes, which are transforming events in prostate cancer, AML and Ewing's sarcoma, associate with PARP1 and DNA-PK and are synthetically lethal with PARP inhibition.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Brenner, J. C. et al. PARP-1 inhibition as a targeted strategy to treat Ewing's sarcoma. Cancer Res. 72, 1608–1613 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Garnett, M. J. et al. Systematic identification of genomic markers of drug sensitivity in cancer cells. Nature 483, 570–575 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Olaussen, K. A. et al. DNA repair by ERCC1 in non-small-cell lung cancer and cisplatin-based adjuvant chemotherapy. New Engl. J. Med. 355, 983–991 (2006).

    Article  CAS  PubMed  Google Scholar 

  126. Moeller, B. J. et al. DNA repair biomarker profiling of head and neck cancer: Ku80 expression predicts locoregional failure and death following radiotherapy. Clin. Can. Res. 17, 2035–2043 (2011).

    Article  CAS  Google Scholar 

  127. Byrski, T. et al. Response to neoadjuvant therapy with cisplatin in BRCA1-positive breast cancer patients. Breast Cancer Res. Treat. 115, 359–363 (2009).

    Article  CAS  PubMed  Google Scholar 

  128. Esteller, M. et al. Promoter hypermethylation and BRCA1 inactivation in sporadic breast and ovarian tumors. J. Natl Cancer Inst. 92, 564–569 (2000). Demonstration that BRCA1 may be silenced in non-hereditary breast and ovarian cancers.

    Article  CAS  PubMed  Google Scholar 

  129. Jazaeri, A. A. et al. Gene expression profiles of BRCA1-linked, BRCA2-linked, and sporadic ovarian cancers. J. Natl Cancer Inst. 94, 990–1000 (2002).

    Article  CAS  PubMed  Google Scholar 

  130. Konstantinopoulos, P. A. et al. Gene expression profile of BRCAness that correlates with responsiveness to chemotherapy and with outcome in patients with epithelial ovarian cancer. J. Clin. Oncol. 28, 3555–3561 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Ahluwalia, A., Hurteau, J. A., Bigsby, R. M. & Nephew, K. P. DNA methylation in ovarian cancer. II. Expression of DNA methyltransferases in ovarian cancer cell lines and normal ovarian epithelial cells. Gynecol. Oncol. 82, 299–304 (2001).

    Article  CAS  PubMed  Google Scholar 

  132. Vollebergh, M. A. Jonkers, J. & Linn, S. C. Genomic instability in breast and ovarian cancers: translation into clinical predictive biomarkers. Cell. Mol. Life Sci. 69, 223–245 (2012).

    Article  CAS  PubMed  Google Scholar 

  133. Willers, H. et al. Utility of DNA repair protein foci for the detection of putative BRCA1-pathway defects in breast cancer biopsies. Mol. Cancer Res. 7, 1304–1309 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Gaymes, T. J. et al. Inhibitors of poly ADP-ribose polymerase (PARP) induce apoptosis of myeloid leukemic cells: potential for therapy of myeloid leukemia and myelodysplastic syndromes. Haematologica 94, 638–646 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Mukhopadhyay, A. et al. Development of a functional assay for homologous recombination status in primary cultures of epithelial ovarian tumor and correlation with sensitivity to poly(ADP-ribose) polymerase inhibitors. Clin. Cancer Res. 16, 2344–2351 (2010).

    Article  CAS  PubMed  Google Scholar 

  136. Graeser, M. et al. A marker of homologous recombination predicts pathologic complete response to neoadjuvant chemotherapy in primary breast cancer. Clin. Cancer Res. 16, 6159–6168 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Pantel, K. & Alix-Panabières, C. Circulating tumour cells in cancer patients: challenges and perspectives. Trends Mol. Med. 16, 398–406 (2010).

    Article  PubMed  Google Scholar 

  138. Bonner, W. M. et al. γH2AX and cancer. Nature Rev. Cancer 8, 957–967 (2008).

    Article  CAS  Google Scholar 

  139. Tanaka, T., Halicka, H. D., Traganos, F., Seiter, K. & Darzynkiewicz, Z. Induction of ATM activation, histone H2AX phosphorylation and apoptosis by etoposide: relation to cell cycle phase. Cell Cycle. 6, 371–376 (2007).

    Article  CAS  PubMed  Google Scholar 

  140. Redon, C. E. et al. Histone γH2AX and poly(ADP-ribose) as clinical pharmacodynamic biomarkers. Clin. Cancer Res. 16, 4532–4542 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Fong, P. C. et al. Inhibition of poly(ADP-ribose) polymerase in tumors from BRCA mutation carriers. N. Engl. J. Med. 361, 123–134 (2009).

    Article  CAS  PubMed  Google Scholar 

  142. Kummar, S. et al. Phase I study of PARP inhibitor ABT-888 in combination with topotecan in adults with refractory solid tumors and lymphomas. Cancer Res. 71, 5626–5634 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Kummar, S. et al. A phase I study of veliparib in combination with metronomic cyclophosphamide in adults with refractory solid tumors and lymphomas. Clin. Cancer Res. 18, 1726–1734 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Almeida, K. H. & Sobol, R. W. A unified view of base excision repair: lesion-dependent protein complexes regulated by post-translational modification. DNA Repair 6, 695–711 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Kurimasa, A. et al. Requirement for the kinase activity of human DNA-dependent protein kinase catalytic subunit in DNA strand break rejoining. Mol. Cell. Biol. 19, 3877–3884 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Zhong, Q. et al. Association of BRCA1 with the hRad50-hMre11-p95 complex and the DNA damage response. Science 285, 747–770 (1999).

    Article  CAS  PubMed  Google Scholar 

  147. Haince, J. F. et al. PARP1-dependent kinetics of recruitment of MRE11 and NBS1 proteins to multiple DNA damage sites. J. Biol. Chem. 283, 1197–1208 (2008).

    Article  CAS  PubMed  Google Scholar 

  148. Stiff, T. et al. ATM and DNA-PK function redundantly to phosphorylate H2AX after exposure to ionizing radiation. Cancer Res. 64, 2390–2396 (2004).

    Article  CAS  PubMed  Google Scholar 

  149. Jensen, R. B., Carreira, A. & Kowalczykowski, S. C. Purified human BRCA2 stimulates RAD51-mediated recombination. Nature 467, 678–683 (2010). The first successful purification of full-length BRCA2, demonstrating that it binds RAD51 and promotes the displacement of replication protein A (RPA) on the single-stranded DNA filament that invades the complementary sequence in the sister chromatid.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Liu, L. & Gerson, S. L. Targeted modulation of MGMT: clinical implications. Clin. Cancer Res. 12, 328–331 (2006).

    Article  CAS  PubMed  Google Scholar 

  151. Jaeckle, K. A. et al. Correlation of tumor O6 methylguanine-DNA methyltransferase levels with survival of malignant astrocytoma patients treated with bis-chloroethylnitrosourea: a Southwest Oncology Group study. J. Clin. Oncol. 16, 3310–3315 (1998).

    Article  CAS  PubMed  Google Scholar 

  152. Zhang, X. et al. Polymorphisms in DNA base excision repair genes ADPRT and XRCC1 and risk of lung cancer. Cancer Res. 65, 722–726 (2005).

    CAS  PubMed  Google Scholar 

  153. Zaremba, T. et al. Poly(ADP-ribose) polymerase-1 (PARP-1) parmacogenetics, activity and expression analysis in cancer patients and healthy volunteers. Biochem. J. 436, 671–679 (2011).

    Article  CAS  PubMed  Google Scholar 

  154. Jalal, S., Earley, J. N. & Turchi, J. J. DNA repair: from genome maintenance to biomarker and therapeutic target. Clin. Cancer Res. 17, 6973–6984 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. di Masi, A. & Antoccia, A. NBS1 Heterozygosity and cancer risk. Curr. Genom. 9, 275–281 (2008).

    Article  CAS  Google Scholar 

  156. Digweed, M. & Sperling, K. Nijmegen breakage syndrome: clinical manifestation of defective response to DNA double-strand breaks. DNA Repair 3, 1207–1217 (2004).

    Article  CAS  PubMed  Google Scholar 

  157. Heikkinen, K., Karppinen, S. M., Soini, Y., Makinen, M. & Winqvist, R. Mutation screening of Mre11 complex genes: indication of RAD50 involvement in breast and ovarian cancer susceptibility. J. Med.Genet. 40, e131 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Giannini, G. et al. Human MRE11 is inactivated in mismatch repair-deficient cancers. EMBO Rep. 3, 248–254 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Takemura, H. et al. Defective Mre11-dependent activation of Chk2 by ataxia telangiectasia mutated in colorectal carcinoma cells in response to replication-dependent DNA double strand breaks. J. Biol. Chem. 281, 30814–30823 (2006).

    Article  CAS  PubMed  Google Scholar 

  160. Willems, P. et al. Polymorphisms in nonhomologous end-joining genes associated with breast cancer risk and chromosomal radiosensitivity. Genes Chromosomes Cancer 47, 137–148 (2008).

    Article  CAS  PubMed  Google Scholar 

  161. Lee, M. N. et al. Epigenetic inactivation of the chromosomal stability control genes BRCA1, BRCA2, and XRCC5 in non-small cell lung cancer. Clin. Cancer Res. 13, 832–838 (2007).

    Article  CAS  PubMed  Google Scholar 

  162. Liu, Y. et al. Polymorphisms of LIG4 and XRCC4 involved in the NHEJ pathway interact to modify risk of glioma. Hum. Mutat. 29, 381–389 (2008).

    Article  PubMed  CAS  Google Scholar 

  163. Yan, D. et al. Targeting DNA-PKcs and ATM with miR-101 sensitizes tumors to radiation. PLoS ONE 5, e11397 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  164. Berman, D. B. et al. A common mutation in BRCA2 that predisposes to a variety of cancers is found in both Jewish Ashkenazi and non-Jewish individuals. Cancer Res. 56, 3409–3414 (1996).

    CAS  PubMed  Google Scholar 

  165. Brose, M. S. et al. Cancer risk estimates for BRCA1 mutation carriers identified in a risk evaluation program. J. Natl Cancer Inst. 94, 1365–1372 (2002).

    Article  CAS  PubMed  Google Scholar 

  166. Mohindra, A. et al. Defects in homologous recombination repair in mismatch-repair-deficient tumour cell lines. Hum. Mol. Genet. 11, 2189–2200 (2002).

    Article  CAS  PubMed  Google Scholar 

  167. Kim, N. G. et al. Frameshift mutations at coding mononucleotide repeats of the hRAD50 gene in gastrointestinal carcinomas with microsatellite instability. Cancer Res. 61, 36–38 (2001).

    Article  CAS  PubMed  Google Scholar 

  168. Rosenberg, P. S., Greene, M. H. & Alter, B. P. Cancer incidence in persons with Fanconi anemia. Blood 101, 822–826 (2003).

    Article  CAS  PubMed  Google Scholar 

  169. Marsit, C. J. et al. Inactivation of the Fanconi anemia/BRCA pathway in lung and oral cancers: implications for treatment and survival. Oncogene 23, 1000–1004 (2004).

    Article  CAS  PubMed  Google Scholar 

  170. Thompson, D. et al. Cancer risks and mortality in heterozygous ATM mutation carriers. J. Natl Cancer Inst. 97, 813–822 (2005). An epidemiological study of the cancer incidence and mortality of >1,000 individuals with heterozygous ATM mutations, identifying an increased risk of breast cancer and some other cancer types.

    Article  CAS  PubMed  Google Scholar 

  171. Ai, L. et al. Ataxia-telangiectasia-mutated (ATM) gene in head and neck squamous cell carcinoma: promoter hypermethylation with clinical correlation in 100 cases. Cancer Epidemiol. Biomarkers Prev. 13, 150–156 (2004).

    Article  CAS  PubMed  Google Scholar 

  172. Antoni, L., Sodha, N., Collins, I. & Garrett, M. D. CHK2 kinase: cancer susceptibility and cancer therapy - two sides of the same coin? Nature Rev. Cancer 7, 925–936 (2007).

    Article  CAS  Google Scholar 

  173. Panda, H. et al. Amino acid Asp181 of 5′-flap endonuclease 1 is a useful target for chemotherapeutic development. Biochemistry 48, 9952–9958 (2009).

    Article  CAS  PubMed  Google Scholar 

  174. Gao, Z., Maloney, D. J., Dedkova, L. M. & Hecht, S. M. Inhibitors of DNA polymerase β: activity and mechanism. Bioorg. Med. Chem. 16, 4331–4340 (2008).

    Article  CAS  PubMed  Google Scholar 

  175. Kumamoto-Yonezawa, Y. et al. Enhancement of human cancer cell radiosensitivity by conjugated eicosapentaenoic acid - a mammalian DNA polymerase inhibitor. Int. J. Oncol. 36, 577–584 (2010).

    CAS  PubMed  Google Scholar 

  176. Chen, X. et al. Rational design of human DNA ligase inhibitors that target cellular DNA replication and repair. Cancer Res. 68, 3169–3177 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Weiss, G. J. et al. Final results from a phase 1 study of oral TRC102 (methoxyamine HCl), an inhibitor of base-excision repair, to potentiate the activity of pemetrexedin patients with refractory cancer. J. Clin. Oncol. Abstr. 2576 (2010).

  178. Madhusudan, S. et al. Isolation of a small molecule inhibitor of DNA base excision repair. Nucleic Acids Res. 33, 4711–4724 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Tutt, A. et al. Oral poly(ADP-ribose) polymerase inhibitor olaparib in patients with BRCA1 or BRCA2 mutations and advanced breast cancer: a proof-of-concept trial. Lancet 376, 235–244 (2010).

    Article  CAS  PubMed  Google Scholar 

  180. Samol, J. et al. Safety and tolerability of the poly(ADP-ribose) polymerase (PARP) inhibitor, olaparib (AZD2281) in combination with topotecan for the treatment of patients with advanced solid tumors: a phase I study. Invest. New Drugs 30, 1493–1500 (2012).

    Article  CAS  PubMed  Google Scholar 

  181. Khan, O. A. et al. A phase I study of the safety and tolerability of olaparib (AZD2281, KU0059436) and dacarbazine in patients with advanced solid tumours. Br. J. Cancer 104, 750–755 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Patel, A. G., De Lorenzo, S., Flatten, K., Poirier, G. & Kaufmann, S. H. Failure of iniparib to inhibit poly(ADP-ribose) polymerase in vitro. Clin. Cancer Res. 18, 1655–1662 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Kuschal, C. et al. Cyclosporin A inhibits nucleotide excision repair via downregulation of the xeroderma pigmentosum group A and G proteins, which is mediated by calcineurin inhibition. Exp. Dermatol. 20, 795–799 (2011).

    Article  CAS  PubMed  Google Scholar 

  184. Prewett, M. et al. Tumors established with cell lines selected for oxaliplatin resistance respond to oxaliplatin if combined with cetuximab. Clin. Cancer Res. 13, 7432–7440 (2007).

    Article  CAS  PubMed  Google Scholar 

  185. Kruszewski, M., Wojewodzka, M., Iwanenko, T., Szumiel, I. & Okuyama, A. Differential inhibitory effect of OK-1035 on DNA repair in L5178Y murine lymphoma sublines with functional or defective repair of double strand breaks. Mutat. Res. 409, 31–36 (1998).

    Article  CAS  PubMed  Google Scholar 

  186. Ismail, I. H. et al. SU11752 inhibits the DNA-dependent protein kinase and DNA double-strand break repair resulting in ionizing radiation sensitization. Oncogene 23, 873–882 (2004).

    Article  CAS  PubMed  Google Scholar 

  187. Huang, F. et al. Identification of specific inhibitors of human RAD51 recombinase using high-throughput screening. ACS Chem. Biol. 6, 628–635 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Rainey, M. D., Charlton, M. E., Stanton, R. V. & Kastan, M. B. Transient inhibition of ATM kinase is sufficient to enhance cellular sensitivity to ionizing radiation. Cancer Res. 68, 7466–7467 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Golding, S. E. et al. Improved ATM kinase inhibitor KU-60019 radiosensitizes glioma cells, compromises insulin, AKT and ERK prosurvival signaling, and inhibits migration and invasion. Mol. Cancer Ther. 8, 2894–2902 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Hirai, H. et al. Small-molecule inhibition of Wee1 kinase by MK-1775 selectively sensitizes p53-deficient tumor cells to DNA-damaging agents. Mol. Cancer Ther. 8, 2992–3000 (2009).

    Article  CAS  PubMed  Google Scholar 

  191. Rajeshkumar, N. V. et al. MK-1775, a potent Wee1 inhibitor, synergizes with gemcitabine to achieve tumor regressions, selectively in p53-deficient pancreatic cancer xenografts. Clin. Cancer Res. 17, 2799–2806 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Kreahling, J. M. et al. MK1775, a selective Wee1 inhibitor, shows single-agent antitumor activity against sarcoma cells. Mol. Cancer Ther. 11, 174–182 (2012).

    Article  CAS  PubMed  Google Scholar 

  193. Schellens, J. H. M. et al. Update on a phase I pharmacologic and pharmacodynamic study of MK-1775, a Wee1 tyrosine kinase inhibitor, in monotherapy and combination with gemcitabine, cisplatin, or carboplatin in patients with advanced solid tumors. J. Clin. Oncol. Abstr. 29, 3068 (2011).

    Article  Google Scholar 

  194. Brezak, M. C. et al. IRC-083864, a novel bisquinone inhibitor of CDC25 phosphatases active against human cancer cells. Int. J. Cancer 124, 1449–1456 (2009).

    Article  CAS  PubMed  Google Scholar 

  195. Lavecchia, A., Di Giovanni, C. & Novellino, E. Inhibitors of Cdc25 phosphatases as anticancer agents: a patent review. Expert Opin. Ther. Patents 20, 405–425 (2010).

    Article  CAS  Google Scholar 

  196. Jobson, A. G. et al. Cellular inhibition of checkpoint kinase 2 (Chk2) and potentiation of camptothecins and radiation by the novel Chk2 inhibitor PV1019 [7-nitro-1H-indole-2-carboxylic acid {4-[1-(guanidinohydrazone)-ethyl]-phenyl}-amide]. J. Pharmacol. Exp. Ther. 331, 816–826 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Swisher, E. M. et al. Secondary Brca1 mutations in Brca1-mutated ovarian carcinomas with platinum resistance. Cancer Res. 68, 2581–2586 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The author would like to thank her laboratory group and colleagues for their help and support and CR UK, Agouron (then Pfizer, then Clovis), KuDOS, AstraZeneca, EPSRC and MRC for supporting her research. Special thanks go to Z. Hostomsky, who was her PARP champion at Agouron and Pfizer.

Author information

Authors and Affiliations

Authors

Ethics declarations

Competing interests

N.J.C. is a named inventor on PARP inhibitor and DNA-PK inhibitor patents. N.J.C. is currently or previously in receipt of research funding from Agouron Pharmaceuticals, BioMarin, BiPAR and Pfizer for PARP-related studies and from AstraZeneca and KuDOS for DNA-PK and ATM-related studies and from Vertex for ATR-related studies. N.J.C. has received honoraria for speaking at AACR, ASCO, Abbott and Consultancy funding from BioMarin and Eisai.

Related links

Glossary

Cell cycle checkpoints

Points in the cell division cycle at which the cell may arrest in response to stress; for example, following DNA damage to prevent damage becoming inherited by daughter cells. The G1 checkpoint controls entry into S phase, the S phase checkpoint halts the progression of S phase and the G2 checkpoint controls entry into mitosis.

Replication stress

The harmful effect of partially replicated DNA. It can be caused by oncogene-induced hyper-replication that activates origins more than once per S phase, by nucleotide pool imbalance or by DNA damage; for example, by reactive oxygen species.

Synthetic lethality

Cell death caused by inactivation (mutation or inhibition) of two genes or their products or two pathways, when inactivation of either alone is not lethal.

Alkylations

Transfers of an alkyl group from one molecule to another; for example, the transfer of a methyl group from temozolomide to guanine.

Myelosuppression

Impairment of bone marrow function resulting in reduced numbers of blood cells (red or white) or platelets.

Phase III trials

Test the efficacy of the novel drug, or combination, and side effects compared with the standard-of-care for that tumour type, with patients randomized to two groups. Numbers of patients are larger (1,000–3,000) than other Phase trials.

Phase II trials

Use the maximum tolerated dose to treat groups of patients with a particular tumour type, these studies determine whether the drug, or drug combination, is effective in one or more tumour types and are used to further monitor drug safety. These generally involve small numbers of patients (100–300).

Deamination

Removal of an amide group; for example, deamination of adenosine to inosine (adenine to hypoxanthine).

Oxidative phosphorylation

Metabolic pathway for the generation of ATP occurring in the mitochondria of eukaryotes in which NADH and succinate from the Krebs' cycle are oxidized by the electron transport chain. The process also generates reactive oxygen species.

Abasic sites

(Also known as apurinic or apyrimidinic (AP) sites). A site on the DNA where there is no purine or pyrimidine base.

Antimetabolites

Substances that resemble a normal precursor or cofactor, usually for DNA synthesis, which interfere with the normal metabolic process; for example, nucleoside analogues such as 6-thioguanine or folate analogues such as pemetrexed.

Phase I trial

Determines pharmacokinetics (absorption, distribution, metabolism and excretion (ADME)) and the safe dose or maximum tolerated dose of a novel agent or combination. They are conducted in 20–80 patients with a variety of tumour types who have often undergone treatment with standard-of-care but for whom no effective therapy options are available. These studies usually involve escalating the dose of the novel agent in successive cohorts of patients until unacceptable toxicities are seen.

Microsatellite instability

(MSI). Microsatellites are sequences in DNA; for example, ten thymidines in direct sequence. MSI occurs when this sequence is shortened or lengthened owing to defects in the mismatch repair system.

Radiomimetics

Drugs that introduce the same DNA damage as ionizing radiation.

Fanconi anaemia

A rare genetic disorder that results in aplastic anaemia, leukaemia and cancer susceptibility, and hypersensitivity to DNA crosslinking agents. The pathway is responsible for the repair of DNA interstrand crosslinks and overlaps somewhat with homologous recombination repair.

Crosslinking agents

Molecules with two reactive groups (for example, bifunctional alkylating agents and cisplatin) that can react with two groups in DNA on the same strand to form intrastrand crosslinks or opposite strands to form interstrand crosslinks.

Focus

The accumulation of a substance (usually protein) that may be identified and visualized (usually by a fluorescently tagged antibody) in one spot.

Pharmacodynamic

The physiological or biochemical effect of a drug on the body. A pharmacodynamic biomarker is a measure of this effect; for example, the product of an enzyme reaction may be reduced by a drug that inhibits the enzyme.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Curtin, N. DNA repair dysregulation from cancer driver to therapeutic target. Nat Rev Cancer 12, 801–817 (2012). https://doi.org/10.1038/nrc3399

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrc3399

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