Skip to main content

Advertisement

Log in

Targeting DNA damage response pathways in cancer

  • Review Article
  • Published:

From Nature Reviews Cancer

View current issue Sign up to alerts

Abstract

Cells have evolved a complex network of biochemical pathways, collectively known as the DNA damage response (DDR), to prevent detrimental mutations from being passed on to their progeny. The DDR coordinates DNA repair with cell-cycle checkpoint activation and other global cellular responses. Genes encoding DDR factors are frequently mutated in cancer, causing genomic instability, an intrinsic feature of many tumours that underlies their ability to grow, metastasize and respond to treatments that inflict DNA damage (such as radiotherapy). One instance where we have greater insight into how genetic DDR abrogation impacts on therapy responses is in tumours with mutated BRCA1 or BRCA2. Due to compromised homologous recombination DNA repair, these tumours rely on alternative repair mechanisms and are susceptible to chemical inhibitors of poly(ADP-ribose) polymerase (PARP), which specifically kill homologous recombination-deficient cancer cells, and have become a paradigm for targeted cancer therapy. It is now clear that many other synthetic-lethal relationships exist between DDR genes. Crucially, some of these interactions could be exploited in the clinic to target tumours that become resistant to PARP inhibition. In this Review, we discuss state-of-the-art strategies for DDR inactivation using small-molecule inhibitors and highlight those compounds currently being evaluated in the clinic.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1: Cell cycle-dependent DDR activation at DSBs.
Fig. 2: Major and backup pathways of DSB repair.

Similar content being viewed by others

References

  1. Jackson, S. P. & Bartek, J. The DNA-damage response in human biology and disease. Nature 461, 1071–1078 (2009).

    Article  CAS  Google Scholar 

  2. Taylor, A. M. R. et al. Chromosome instability syndromes. Nat. Rev. Dis. Prim. 5, 64 (2019).

    Article  Google Scholar 

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

    Article  CAS  Google Scholar 

  4. Lord, C. J. & Ashworth, A. PARP inhibitors: synthetic lethality in the clinic. Science 355, 1152–1158 (2017).

    Article  CAS  Google Scholar 

  5. 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  Google Scholar 

  6. Farmer, H. et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 434, 917–921 (2005). Along with Bryant et al. (2005), this article shows that PARP inhibition targets BRCA1/2-deficient cells.

    Article  CAS  Google Scholar 

  7. Fugger, K., Hewitt, G., West, S. C. & Boulton, S. J. Tackling PARP inhibitor resistance. Trends Cancer 7, 1102–1118 (2021).

    Article  CAS  Google Scholar 

  8. Blackford, A. N. & Jackson, S. P. ATM, ATR, and DNA-PK: the trinity at the heart of the DNA damage response. Mol. Cell 66, 801–817 (2017).

    Article  CAS  Google Scholar 

  9. Matsuoka, S. et al. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 316, 1160–1166 (2007).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  11. Blackford, A. N. & Stucki, M. How cells respond to DNA breaks in mitosis. Trends Biochem. Sci. 45, 321–331 (2020).

    Article  CAS  Google Scholar 

  12. Leimbacher, P. A. et al. MDC1 interacts with TOPBP1 to maintain chromosomal stability during mitosis. Mol. Cell 74, 571–583 e578 (2019).

    Article  CAS  Google Scholar 

  13. De Marco Zompit, M. et al. The CIP2A-TOPBP1 complex safeguards chromosomal stability during mitosis. Nat. Commun. 13, 4143 (2022).

    Article  Google Scholar 

  14. Mailand, N. et al. Rapid destruction of human Cdc25A in response to DNA damage. Science 288, 1425–1429 (2000).

    Article  CAS  Google Scholar 

  15. Peng, C. Y. et al. Mitotic and G2 checkpoint control: regulation of 14-3-3 protein binding by phosphorylation of Cdc25C on serine-216. Science 277, 1501–1505 (1997).

    Article  CAS  Google Scholar 

  16. Sanchez, Y. et al. Conservation of the Chk1 checkpoint pathway in mammals: linkage of DNA damage to Cdk regulation through Cdc25. Science 277, 1497–1501 (1997).

    Article  CAS  Google Scholar 

  17. Heald, R., McLoughlin, M. & McKeon, F. Human wee1 maintains mitotic timing by protecting the nucleus from cytoplasmically activated Cdc2 kinase. Cell 74, 463–474 (1993).

    Article  CAS  Google Scholar 

  18. Gu, Y., Rosenblatt, J. & Morgan, D. O. Cell cycle regulation of CDK2 activity by phosphorylation of Thr160 and Tyr15. EMBO J. 11, 3995–4005 (1992).

    Article  CAS  Google Scholar 

  19. Booher, R. N., Holman, P. S. & Fattaey, A. Human Myt1 is a cell cycle-regulated kinase that inhibits Cdc2 but not Cdk2 activity. J. Biol. Chem. 272, 22300–22306 (1997).

    Article  CAS  Google Scholar 

  20. Liu, F., Stanton, J. J., Wu, Z. & Piwnica-Worms, H. The human Myt1 kinase preferentially phosphorylates Cdc2 on threonine 14 and localizes to the endoplasmic reticulum and Golgi complex. Mol. Cell Biol. 17, 571–583 (1997).

    Article  CAS  Google Scholar 

  21. Simoneau, A. & Zou, L. An extending ATR-CHK1 circuitry: the replication stress response and beyond. Curr. Opin. Genet. Dev. 71, 92–98 (2021).

    Article  CAS  Google Scholar 

  22. Lecona, E. & Fernandez-Capetillo, O. Targeting ATR in cancer. Nat. Rev. Cancer 18, 586–595 (2018).

    Article  CAS  Google Scholar 

  23. Dupre, A., Boyer-Chatenet, L. & Gautier, J. Two-step activation of ATM by DNA and the Mre11-Rad50-Nbs1 complex. Nat. Struct. Mol. Biol. 13, 451–457 (2006).

    Article  CAS  Google Scholar 

  24. Lee, J. H. & Paull, T. T. ATM activation by DNA double-strand breaks through the Mre11-Rad50-Nbs1 complex. Science 308, 551–554 (2005).

    Article  CAS  Google Scholar 

  25. Zou, L. & Elledge, S. J. Sensing DNA damage through ATRIP recognition of RPA-ssDNA complexes. Science 300, 1542–1548 (2003).

    Article  CAS  Google Scholar 

  26. Bass, T. E. et al. ETAA1 acts at stalled replication forks to maintain genome integrity. Nat. Cell Biol. 18, 1185–1195 (2016).

    Article  CAS  Google Scholar 

  27. Feng, S. et al. Ewing tumor-associated antigen 1 interacts with replication protein a to promote restart of stalled replication forks. J. Biol. Chem. 291, 21956–21962 (2016).

    Article  CAS  Google Scholar 

  28. Haahr, P. et al. Activation of the ATR kinase by the RPA-binding protein ETAA1. Nat. Cell Biol. 18, 1196–1207 (2016).

    Article  CAS  Google Scholar 

  29. Kumagai, A., Lee, J., Yoo, H. Y. & Dunphy, W. G. TopBP1 activates the ATR-ATRIP complex. Cell 124, 943–955 (2006).

    Article  CAS  Google Scholar 

  30. Caldecott, K. W. DNA single-strand break repair. Exp. Cell Res. 329, 2–8 (2014).

    Article  CAS  Google Scholar 

  31. Jiricny, J. Postreplicative mismatch repair. Cold Spring Harb. Perspect. Biol. 5, a012633 (2013).

    Article  Google Scholar 

  32. Klein, H. L. Genome instabilities arising from ribonucleotides in DNA. DNA Repair 56, 26–32 (2017).

    Article  CAS  Google Scholar 

  33. Semlow, D. R. & Walter, J. C. Mechanisms of vertebrate DNA interstrand cross-link repair. Annu. Rev. Biochem. 90, 107–135 (2021).

    Article  CAS  Google Scholar 

  34. Ruggiano, A. & Ramadan, K. DNA-protein crosslink proteases in genome stability. Commun. Biol. 4, 11 (2021).

    Article  CAS  Google Scholar 

  35. Davidovic, L., Vodenicharov, M., Affar, E. B. & Poirier, G. G. Importance of poly(ADP-ribose) glycohydrolase in the control of poly(ADP-ribose) metabolism. Exp. Cell Res. 268, 7–13 (2001).

    Article  CAS  Google Scholar 

  36. Tarsounas, M. & Sung, P. The antitumorigenic roles of BRCA1-BARD1 in DNA repair and replication. Nat. Rev. Mol. Cell Biol. 21, 284–299 (2020).

    Article  CAS  Google Scholar 

  37. Spies, M. & Fishel, R. Mismatch repair during homologous and homologous recombination. Cold Spring Harb. Perspect. Biol. 7, a022657 (2015).

    Article  Google Scholar 

  38. Nambiar, T. S., Baudrier, L., Billon, P. & Ciccia, A. CRISPR-based genome editing through the lens of DNA repair. Mol. Cell 82, 348–388 (2022).

    Article  CAS  Google Scholar 

  39. Walker, J. R., Corpina, R. A. & Goldberg, J. Structure of the Ku heterodimer bound to DNA and its implications for double-strand break repair. Nature 412, 607–614 (2001).

    Article  CAS  Google Scholar 

  40. Graham, T. G. W., Walter, J. C. & Loparo, J. J. Two-stage synapsis of DNA ends during non-homologous end joining. Mol. Cell 61, 850–858 (2016).

    Article  CAS  Google Scholar 

  41. Dobbs, T. A., Tainer, J. A. & Lees-Miller, S. P. A structural model for regulation of NHEJ by DNA-PKcs autophosphorylation. DNA Repair 9, 1307–1314 (2010).

    Article  CAS  Google Scholar 

  42. Hammel, M. et al. Ku and DNA-dependent protein kinase dynamic conformations and assembly regulate DNA binding and the initial non-homologous end joining complex. J. Biol. Chem. 285, 1414–1423 (2010).

    Article  Google Scholar 

  43. Critchlow, S. E., Bowater, R. P. & Jackson, S. P. Mammalian DNA double-strand break repair protein XRCC4 interacts with DNA ligase IV. Curr. Biol. 7, 588–598 (1997).

    Article  CAS  Google Scholar 

  44. Grawunder, U. et al. Activity of DNA ligase IV stimulated by complex formation with XRCC4 protein in mammalian cells. Nature 388, 492–495 (1997).

    Article  CAS  Google Scholar 

  45. Ahnesorg, P., Smith, P. & Jackson, S. P. XLF interacts with the XRCC4-DNA ligase IV complex to promote DNA nonhomologous end-joining. Cell 124, 301–313 (2006).

    Article  CAS  Google Scholar 

  46. Balmus, G. et al. Synthetic lethality between PAXX and XLF in mammalian development. Genes Dev. 30, 2152–2157 (2016).

    Article  CAS  Google Scholar 

  47. Lescale, C. et al. Specific roles of XRCC4 paralogs PAXX and XLF during V(D)J recombination. Cell Rep. 16, 2967–2979 (2016).

    Article  CAS  Google Scholar 

  48. Buck, D. et al. Cernunnos, a novel nonhomologous end-joining factor, is mutated in human immunodeficiency with microcephaly. Cell 124, 287–299 (2006).

    Article  CAS  Google Scholar 

  49. Ochi, T. et al. DNA repair. PAXX, a paralog of XRCC4 and XLF, interacts with Ku to promote DNA double-strand break repair. Science 347, 185–188 (2015).

    Article  CAS  Google Scholar 

  50. Zhao, B., Rothenberg, E., Ramsden, D. A. & Lieber, M. R. The molecular basis and disease relevance of non-homologous DNA end joining. Nat. Rev. Mol. Cell Biol. 21, 765–781 (2020).

    Article  CAS  Google Scholar 

  51. Cejka, P. & Symington, L. S. DNA end resection: mechanism and control. Annu. Rev. Genet. 55, 285–307 (2021).

    Article  Google Scholar 

  52. Huertas, P. & Jackson, S. P. Human CtIP mediates cell cycle control of DNA end resection and double strand break repair. J. Biol. Chem. 284, 9558–9565 (2009).

    Article  CAS  Google Scholar 

  53. Sartori, A. A. et al. Human CtIP promotes DNA end resection. Nature 450, 509–514 (2007).

    Article  CAS  Google Scholar 

  54. van Sluis, M. & McStay, B. A localized nucleolar DNA damage response facilitates recruitment of the homology-directed repair machinery independent of cell cycle stage. Genes Dev. 29, 1151–1163 (2015).

    Article  Google Scholar 

  55. Yilmaz, D. et al. Activation of homologous recombination in G1 preserves centromeric integrity. Nature 600, 748–753 (2021).

    Article  CAS  Google Scholar 

  56. Gravel, S., Chapman, J. R., Magill, C. & Jackson, S. P. DNA helicases Sgs1 and BLM promote DNA double-strand break resection. Genes Dev. 22, 2767–2772 (2008).

    Article  CAS  Google Scholar 

  57. Zhu, Z., Chung, W. H., Shim, E. Y., Lee, S. E. & Ira, G. Sgs1 helicase and two nucleases DNA2 and Exo1 resect DNA double-strand break ends. Cell 134, 981–994 (2008).

    Article  CAS  Google Scholar 

  58. Mimitou, E. P. & Symington, L. S. Sae2, Exo1 and Sgs1 collaborate in DNA double-strand break processing. Nature 455, 770–774 (2008).

    Article  CAS  Google Scholar 

  59. West, S. C. et al. Resolution of recombination intermediates: mechanisms and regulation. Cold Spring Harb. Symp. Quant. Biol. 80, 103–109 (2015).

    Article  Google Scholar 

  60. Llorens-Agost, M. et al. POLθ-mediated end joining is restricted by RAD52 and BRCA2 until the onset of mitosis. Nat. Cell Biol. 23, 1095–1104 (2021).

    Article  CAS  Google Scholar 

  61. 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  Google Scholar 

  62. Karanam, K., Kafri, R., Loewer, A. & Lahav, G. Quantitative live cell imaging reveals a gradual shift between DNA repair mechanisms and a maximal use of HR in mid S phase. Mol. Cell 47, 320–329 (2012).

    Article  CAS  Google Scholar 

  63. Pellegrino, S., Michelena, J., Teloni, F., Imhof, R. & Altmeyer, M. Replication-coupled dilution of H4K20me2 guides 53BP1 to pre-replicative chromatin. Cell Rep. 19, 1819–1831 (2017).

    Article  CAS  Google Scholar 

  64. Simonetta, M. et al. H4K20me2 distinguishes pre-replicative from post-replicative chromatin to appropriately direct DNA repair pathway choice by 53BP1-RIF1-MAD2L2. Cell Cycle 17, 124–136 (2018).

    Article  CAS  Google Scholar 

  65. Fradet-Turcotte, A. et al. 53BP1 is a reader of the DNA-damage-induced H2A Lys 15 ubiquitin mark. Nature 499, 50–54 (2013).

    Article  CAS  Google Scholar 

  66. Botuyan, M. V. et al. Structural basis for the methylation state-specific recognition of histone H4-K20 by 53BP1 and Crb2 in DNA repair. Cell 127, 1361–1373 (2006).

    Article  CAS  Google Scholar 

  67. Feng, S. et al. RIF1-ASF1-mediated high-order chromatin structure safeguards genome integrity. Nat. Commun. 13, 957 (2022).

    Article  CAS  Google Scholar 

  68. Mirman, Z., Sasi, N. K., King, A., Chapman, J. R. & de Lange, T. 53BP1-shieldin-dependent DSB processing in BRCA1-deficient cells requires CST-Polα-primase fill-in synthesis. Nat. Cell Biol. 24, 51–61 (2022).

    Article  CAS  Google Scholar 

  69. Ochs, F. et al. Stabilization of chromatin topology safeguards genome integrity. Nature 574, 571–574 (2019).

    Article  CAS  Google Scholar 

  70. Paiano, J. et al. Role of 53BP1 in end protection and DNA synthesis at DNA breaks. Genes Dev. 35, 1356–1367 (2021).

    Article  CAS  Google Scholar 

  71. Balmus, G. et al. ATM orchestrates the DNA-damage response to counter toxic non-homologous end-joining at broken replication forks. Nat. Commun. 10, 87 (2019).

    Article  CAS  Google Scholar 

  72. Becker, J. R. et al. BARD1 reads H2A lysine 15 ubiquitination to direct homologous recombination. Nature 596, 433–437 (2021).

    Article  CAS  Google Scholar 

  73. Hu, Q. et al. Mechanisms of BRCA1-BARD1 nucleosome recognition and ubiquitylation. Nature 596, 438–443 (2021).

    Article  CAS  Google Scholar 

  74. Krais, J. J. et al. RNF168-mediated localization of BARD1 recruits the BRCA1-PALB2 complex to DNA damage. Nat. Commun. 12, 5016 (2021).

    Article  CAS  Google Scholar 

  75. Nakamura, K. et al. H4K20me0 recognition by BRCA1-BARD1 directs homologous recombination to sister chromatids. Nat. Cell Biol. 21, 311–318 (2019). This study shows that BARD1 interacts with H4K20me0, which is found only on nascent chromatin. Therefore, HR is restricted to S phase, where all the H4K20 is unmethylated.

    Article  CAS  Google Scholar 

  76. Ochs, F. et al. 53BP1 fosters fidelity of homology-directed DNA repair. Nat. Struct. Mol. Biol. 23, 714–721 (2016).

    Article  CAS  Google Scholar 

  77. Ramsden, D. A., Carvajal-Garcia, J. & Gupta, G. P. Mechanism, cellular functions and cancer roles of polymerase-theta-mediated DNA end joining. Nat. Rev. Mol. Cell Biol. 23, 125–140 (2021).

    Article  Google Scholar 

  78. Thompson, D. & Easton, D. F., Breast Cancer Linkage Consortium. Cancer incidence in BRCA1 mutation carriers. J. Natl Cancer Inst. 94, 1358–1365 (2002).

    Article  CAS  Google Scholar 

  79. Breast Cancer Linkage Consortium. Cancer risks in BRCA2 mutation carriers. J. Natl Cancer Inst. 91, 1310–1316 (1999).

    Article  Google Scholar 

  80. Schlacher, K., Wu, H. & Jasin, M. A distinct replication fork protection pathway connects Fanconi anemia tumor suppressors to RAD51-BRCA1/2. Cancer Cell 22, 106–116 (2012).

    Article  CAS  Google Scholar 

  81. Schlacher, K. et al. Double-strand break repair-independent role for BRCA2 in blocking stalled replication fork degradation by MRE11. Cell 145, 529–542 (2011).

    Article  CAS  Google Scholar 

  82. Quinet, A., Lemacon, D. & Vindigni, A. Replication fork reversal: players and guardians. Mol. Cell 68, 830–833 (2017).

    Article  CAS  Google Scholar 

  83. Roy, R., Chun, J. & Powell, S. N. BRCA1 and BRCA2: different roles in a common pathway of genome protection. Nat. Rev. Cancer 12, 68–78 (2012).

    Article  CAS  Google Scholar 

  84. Ceccaldi, R. et al. Homologous-recombination-deficient tumours are dependent on Poltheta-mediated repair. Nature 518, 258–262 (2015).

    Article  CAS  Google Scholar 

  85. Zhang, Y. & Jasin, M. An essential role for CtIP in chromosomal translocation formation through an alternative end-joining pathway. Nat. Struct. Mol. Biol. 18, 80–84 (2011).

    Article  CAS  Google Scholar 

  86. Yousefzadeh, M. J. et al. Mechanism of suppression of chromosomal instability by DNA polymerase POLQ. PLoS Genet. 10, e1004654 (2014).

    Article  Google Scholar 

  87. Murai, J. et al. Trapping of PARP1 and PARP2 by clinical PARP inhibitors. Cancer Res. 72, 5588–5599 (2012). This study proposes that the cytotoxicity of PARP inhibitors depends on their trapping activity rather than their catalytic inhibitory activity.

    Article  CAS  Google Scholar 

  88. Zimmermann, M. et al. CRISPR screens identify genomic ribonucleotides as a source of PARP-trapping lesions. Nature 559, 285–289 (2018).

    Article  CAS  Google Scholar 

  89. Blessing, C. et al. The oncogenic helicase ALC1 regulates PARP inhibitor potency by trapping PARP2 at DNA breaks. Mol. Cell 80, 862–875 e866 (2020).

    Article  CAS  Google Scholar 

  90. Hewitt, G. et al. Defective ALC1 nucleosome remodeling confers PARPi sensitization and synthetic lethality with HRD. Mol. Cell 81, 767–783 e711 (2021).

    Article  CAS  Google Scholar 

  91. Juhasz, S. et al. The chromatin remodeler ALC1 underlies resistance to PARP inhibitor treatment. Sci. Adv. https://doi.org/10.1126/sciadv.abb8626 (2020).

    Article  Google Scholar 

  92. Verma, P. et al. ALC1 links chromatin accessibility to PARP inhibitor response in homologous recombination-deficient cells. Nat. Cell Biol. 23, 160–171 (2021).

    Article  CAS  Google Scholar 

  93. Krastev, D. B. et al. The ubiquitin-dependent ATPase p97 removes cytotoxic trapped PARP1 from chromatin. Nat. Cell Biol. 24, 62–73 (2022).

    Article  CAS  Google Scholar 

  94. Vaitsiankova, A. et al. PARP inhibition impedes the maturation of nascent DNA strands during DNA replication. Nat. Struct. Mol. Biol. 29, 329–338 (2022).

    Article  CAS  Google Scholar 

  95. Hanzlikova, H. et al. The importance of poly(ADP-ribose) polymerase as a sensor of unligated Okazaki fragments during DNA replication. Mol. Cell 71, e313 (2018).

    Article  Google Scholar 

  96. Cong, K. et al. Replication gaps are a key determinant of PARP inhibitor synthetic lethality with BRCA deficiency. Mol. Cell 81, 3128–3144 e3127 (2021). This study proposes that the accumulation of ssDNA gaps underlies the cytotoxic activity of PARP inhibitors.

    Article  CAS  Google Scholar 

  97. Schoonen, P. M. et al. Progression through mitosis promotes PARP inhibitor-induced cytotoxicity in homologous recombination-deficient cancer cells. Nat. Commun. 8, 15981 (2017).

    Article  CAS  Google Scholar 

  98. Simoneau, A., Xiong, R. & Zou, L. The trans cell cycle effects of PARP inhibitors underlie their selectivity toward BRCA1/2-deficient cells. Genes. Dev. 35, 1271–1289 (2021).

    Article  Google Scholar 

  99. 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). This study describes the use of a selective ATM inhibitor to potentiate the anticancer activity of DNA-damaging agents.

    Article  CAS  Google Scholar 

  100. Bryant, H. E. & Helleday, T. Inhibition of poly (ADP-ribose) polymerase activates ATM which is required for subsequent homologous recombination repair. Nucleic Acids Res. 34, 1685–1691 (2006).

    Article  CAS  Google Scholar 

  101. 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  Google Scholar 

  102. Cremona, C. A. & Behrens, A. ATM signalling and cancer. Oncogene 33, 3351–3360 (2014).

    Article  CAS  Google Scholar 

  103. Bailey, P. et al. Genomic analyses identify molecular subtypes of pancreatic cancer. Nature 531, 47–52 (2016).

    Article  CAS  Google Scholar 

  104. Bunting, S. F. et al. 53BP1 inhibits homologous recombination in Brca1-deficient cells by blocking resection of DNA breaks. Cell 141, 243–254 (2010). This study reveals the competition between 53BP1 and BRCA1 in DSB repair pathway choice.

    Article  CAS  Google Scholar 

  105. Xu, G. et al. REV7 counteracts DNA double-strand break resection and affects PARP inhibition. Nature 521, 541–544 (2015).

    Article  CAS  Google Scholar 

  106. Blackford, A. N. et al. The DNA translocase activity of FANCM protects stalled replication forks. Hum. Mol. Genet. 21, 2005–2016 (2012).

    Article  CAS  Google Scholar 

  107. 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  Google Scholar 

  108. Cai, M. Y. et al. Cooperation of the ATM and Fanconi anemia/BRCA pathways in double-strand break end resection. Cell Rep. 30, 2402–2415 e2405 (2020).

    Article  CAS  Google Scholar 

  109. Wang, C. et al. Genetic vulnerabilities upon inhibition of DNA damage response. Nucleic Acids Res. 49, 8214–8231 (2021).

    Article  CAS  Google Scholar 

  110. Cancer Genome Atlas Research Network. Integrated genomic analyses of ovarian carcinoma. Nature 474, 609–615 (2011).

    Article  Google Scholar 

  111. Reiman, A. et al. Lymphoid tumours and breast cancer in ataxia telangiectasia; substantial protective effect of residual ATM kinase activity against childhood tumours. Br. J. Cancer 105, 586–591 (2011).

    Article  CAS  Google Scholar 

  112. Choi, M., Kipps, T. & Kurzrock, R. ATM mutations in cancer: therapeutic implications. Mol. Cancer Ther. 15, 1781–1791 (2016).

    Article  CAS  Google Scholar 

  113. Blunt, T. et al. Defective DNA-dependent protein kinase activity is linked to V(D)J recombination and DNA repair defects associated with the murine scid mutation. Cell 80, 813–823 (1995).

    Article  CAS  Google Scholar 

  114. Kirchgessner, C. U. et al. DNA-dependent kinase (p350) as a candidate gene for the murine SCID defect. Science 267, 1178–1183 (1995).

    Article  CAS  Google Scholar 

  115. Peterson, S. R. et al. Loss of the catalytic subunit of the DNA-dependent protein kinase in DNA double-strand-break-repair mutant mammalian cells. Proc. Natl Acad. Sci. USA 92, 3171–3174 (1995).

    Article  CAS  Google Scholar 

  116. Fok, J. H. L. et al. AZD7648 is a potent and selective DNA-PK inhibitor that enhances radiation, chemotherapy and olaparib activity. Nat. Commun. 10, 5065 (2019).

    Article  Google Scholar 

  117. Zhao, Y. et al. Preclinical evaluation of a potent novel DNA-dependent protein kinase inhibitor NU7441. Cancer Res. 66, 5354–5362 (2006).

    Article  CAS  Google Scholar 

  118. Gurley, K. E. & Kemp, C. J. Synthetic lethality between mutation in Atm and DNA-PK(cs) during murine embryogenesis. Curr. Biol. 11, 191–194 (2001).

    Article  CAS  Google Scholar 

  119. Callen, E. et al. Essential role for DNA-PKcs in DNA double-strand break repair and apoptosis in ATM-deficient lymphocytes. Mol. Cell 34, 285–297 (2009).

    Article  CAS  Google Scholar 

  120. Riabinska, A. et al. Therapeutic targeting of a robust non-oncogene addiction to PRKDC in ATM-defective tumors. Sci. Transl. Med. 5, 189ra178 (2013).

    Article  Google Scholar 

  121. 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).

    Article  CAS  Google Scholar 

  122. Nghiem, P., Park, P. K., Kim, Y., 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  Google Scholar 

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

    Article  CAS  Google Scholar 

  124. Charrier, J. D. et al. Discovery of potent and selective inhibitors of ataxia telangiectasia mutated and Rad3 related (ATR) protein kinase as potential anticancer agents. J. Med. Chem. 54, 2320–2330 (2011).

    Article  CAS  Google Scholar 

  125. Foote, K. M. et al. Discovery and characterization of AZD6738, a potent inhibitor of ataxia telangiectasia mutated and Rad3 related (ATR) kinase with application as an anticancer agent. J. Med. Chem. 61, 9889–9907 (2018).

    Article  CAS  Google Scholar 

  126. Szydzik, J. et al. ATR inhibition enables complete tumour regression in ALK-driven NB mouse models. Nat. Commun. 12, 6813 (2021).

    Article  CAS  Google Scholar 

  127. Lloyd, R. L. et al. Loss of Cyclin C or CDK8 provides ATR inhibitor resistance by suppressing transcription-associated replication stress. Nucleic Acids Res. 49, 8665–8683 (2021).

    Article  CAS  Google Scholar 

  128. Kwok, M. et al. ATR inhibition induces synthetic lethality and overcomes chemoresistance in TP53- or ATM-defective chronic lymphocytic leukemia cells. Blood 127, 582–595 (2016).

    Article  CAS  Google Scholar 

  129. Reaper, P. M. et al. Selective killing of ATM- or p53-deficient cancer cells through inhibition of ATR. Nat. Chem. Biol. 7, 428–430 (2011). This study describes the first ATR inhibitor (VE-821) and shows that it can kill ATM-deficient tumours.

    Article  CAS  Google Scholar 

  130. Tu, X. et al. ATR inhibition is a promising radiosensitizing strategy for triple-negative breast cancer. Mol. Cancer Ther. 17, 2462–2472 (2018).

    Article  CAS  Google Scholar 

  131. Dillon, M. T. et al. ATR inhibition potentiates the radiation-induced inflammatory tumor microenvironment. Clin. Cancer Res. 25, 3392–3403 (2019).

    Article  CAS  Google Scholar 

  132. Feng, X. et al. ATR inhibition potentiates ionizing radiation-induced interferon response via cytosolic nucleic acid-sensing pathways. EMBO J. 39, e104036 (2020).

    Article  CAS  Google Scholar 

  133. Tang, Z. et al. ATR inhibition induces CDK1-SPOP signaling and enhances anti-PD-L1 cytotoxicity in prostate cancer. Clin. Cancer Res. 27, 4898–4909 (2021).

    Article  CAS  Google Scholar 

  134. Thomas, A. et al. Therapeutic targeting of ATR yields durable regressions in small cell lung cancers with high replication stress. Cancer Cell 39, 566–579.e567 (2021).

    Article  CAS  Google Scholar 

  135. El Touny, L. H. et al. ATR inhibition reverses the resistance of homologous recombination deficient MGMTlow/MMRproficient cancer cells to temozolomide. Oncotarget 12, 2114–2130 (2021).

    Article  Google Scholar 

  136. Wallez, Y. et al. The ATR inhibitor AZD6738 synergizes with gemcitabine in vitro and in vivo to induce pancreatic ductal adenocarcinoma regression. Mol. Cancer Ther. 17, 1670–1682 (2018).

    Article  CAS  Google Scholar 

  137. 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  Google Scholar 

  138. Kim, H. et al. Targeting the ATR/CHK1 axis with PARP inhibition results in tumor regression in BRCA-mutant ovarian cancer models. Clin. Cancer Res. 23, 3097–3108 (2017).

    Article  CAS  Google Scholar 

  139. Schoonen, P. M. et al. Premature mitotic entry induced by ATR inhibition potentiates olaparib inhibition-mediated genomic instability, inflammatory signaling, and cytotoxicity in BRCA2-deficient cancer cells. Mol. Oncol. 13, 2422–2440 (2019).

    Article  CAS  Google Scholar 

  140. Gralewska, P. et al. PARP Inhibition increases the reliance on ATR/CHK1 checkpoint signaling leading to synthetic lethality-an alternative treatment strategy for epithelial ovarian cancer cells independent from HR effectiveness. Int. J. Mol. Sci. https://doi.org/10.3390/ijms21249715 (2020).

    Article  Google Scholar 

  141. Nam, A. R. et al. ATR inhibition amplifies antitumor effects of olaparib in biliary tract cancer. Cancer Lett. 516, 38–47 (2021).

    Article  CAS  Google Scholar 

  142. Yazinski, S. A. et al. ATR inhibition disrupts rewired homologous recombination and fork protection pathways in PARP inhibitor-resistant BRCA-deficient cancer cells. Genes Dev. 31, 318–332 (2017).

    Article  CAS  Google Scholar 

  143. Murai, J. et al. Resistance to PARP inhibitors by SLFN11 inactivation can be overcome by ATR inhibition. Oncotarget 7, 76534–76550 (2016).

    Article  Google Scholar 

  144. 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 (2012).

    Article  CAS  Google Scholar 

  145. Brooks, K. et al. A potent Chk1 inhibitor is selectively cytotoxic in melanomas with high levels of replicative stress. Oncogene 32, 788–796 (2013).

    Article  CAS  Google Scholar 

  146. Massey, A. J. et al. mTORC1 and DNA-PKcs as novel molecular determinants of sensitivity to Chk1 inhibition. Mol. Oncol. 10, 101–112 (2016).

    Article  CAS  Google Scholar 

  147. Ma, C. X. et al. Targeting Chk1 in p53-deficient triple-negative breast cancer is therapeutically beneficial in human-in-mouse tumor models. J. Clin. Invest. 122, 1541–1552 (2012).

    Article  CAS  Google Scholar 

  148. Booth, L. et al. PARP and CHK inhibitors interact to cause DNA damage and cell death in mammary carcinoma cells. Cancer Biol. Ther. 14, 458–465 (2013).

    Article  CAS  Google Scholar 

  149. Sen, T. et al. Targeting DNA damage response promotes antitumor immunity through STING-mediated T-cell activation in small cell lung cancer. Cancer Discov. 9, 646–661 (2019).

    Article  CAS  Google Scholar 

  150. Duda, H. et al. A mechanism for controlled breakage of under-replicated chromosomes during mitosis. Dev. Cell 39, 740–755 (2016).

    Article  CAS  Google Scholar 

  151. Pfister, S. X. et al. Inhibiting WEE1 selectively kills histone H3K36me3-deficient cancers by dNTP starvation. Cancer Cell 28, 557–568 (2015).

    Article  CAS  Google Scholar 

  152. ClinicalTrials.gov https://ClinicalTrials.gov/show/NCT03284385 (2017).

  153. Gallo, D. et al. CCNE1 amplification is synthetic lethal with PKMYT1 kinase inhibition. Nature 604, 749–756 (2022). This study describes the development and use of a selective PKMYT1 inhibitor to target cyclin E1-overexpressing cancer cells.

    Article  CAS  Google Scholar 

  154. Roerink, S. F., van Schendel, R. & Tijsterman, M. Polymerase theta-mediated end joining of replication-associated DNA breaks in C. elegans. Genome Res. 24, 954–962 (2014).

    Article  CAS  Google Scholar 

  155. Mateos-Gomez, P. A. et al. The helicase domain of Poltheta counteracts RPA to promote alt-NHEJ. Nat. Struct. Mol. Biol. 24, 1116–1123 (2017).

    Article  CAS  Google Scholar 

  156. Lu, G. et al. Ligase I and ligase III mediate the DNA double-strand break ligation in alternative end-joining. Proc. Natl Acad. Sci. USA 113, 1256–1260 (2016).

    Article  CAS  Google Scholar 

  157. Kent, T., Chandramouly, G., McDevitt, S. M., Ozdemir, A. Y. & Pomerantz, R. T. Mechanism of microhomology-mediated end-joining promoted by human DNA polymerase theta. Nat. Struct. Mol. Biol. 22, 230–237 (2015).

    Article  CAS  Google Scholar 

  158. Wyatt, D. W. et al. Essential roles for polymerase theta-mediated end joining in the repair of chromosome breaks. Mol. Cell 63, 662–673 (2016).

    Article  CAS  Google Scholar 

  159. Higgins, G. S. et al. A small interfering RNA screen of genes involved in DNA repair identifies tumor-specific radiosensitization by POLQ knockdown. Cancer Res. 70, 2984–2993 (2010).

    Article  CAS  Google Scholar 

  160. Mateos-Gomez, P. A. et al. Mammalian polymerase theta promotes alternative NHEJ and suppresses recombination. Nature 518, 254–257 (2015).

    Article  CAS  Google Scholar 

  161. Shima, N., Munroe, R. J. & Schimenti, J. C. The mouse genomic instability mutation chaos1 is an allele of Polq that exhibits genetic interaction with Atm. Mol. Cell Biol. 24, 10381–10389 (2004).

    Article  CAS  Google Scholar 

  162. Zhou, J. et al. A first-in-class polymerase theta inhibitor selectively targets homologous-recombination-deficient tumors. Nat. Cancer 2, 598–610 (2021).

    Article  CAS  Google Scholar 

  163. Zatreanu, D. et al. Poltheta inhibitors elicit BRCA-gene synthetic lethality and target PARP inhibitor resistance. Nat. Commun. 12, 3636 (2021). Along with Zhou et al. (2021), this article describes the cytotoxicity of a POLQ inhibitor with regard to cancer cells that have acquired resistance to PARP inhibitors.

    Article  CAS  Google Scholar 

  164. Nijman, S. M. et al. The deubiquitinating enzyme USP1 regulates the Fanconi anemia pathway. Mol. Cell 17, 331–339 (2005).

    Article  CAS  Google Scholar 

  165. Huang, T. T. et al. Regulation of monoubiquitinated PCNA by DUB autocleavage. Nat. Cell Biol. 8, 339–347 (2006).

    Article  CAS  Google Scholar 

  166. Cohn, M. A. et al. A UAF1-containing multisubunit protein complex regulates the Fanconi anemia pathway. Mol. Cell 28, 786–797 (2007).

    Article  CAS  Google Scholar 

  167. Lim, K. S. et al. USP1 is required for replication fork protection in BRCA1-deficient tumors. Mol. Cell 72, 925–941 e924 (2018).

    Article  CAS  Google Scholar 

  168. Ulrich, H. D. & Takahashi, T. Readers of PCNA modifications. Chromosoma 122, 259–274 (2013).

    Article  CAS  Google Scholar 

  169. Mailand, N., Gibbs-Seymour, I. & Bekker-Jensen, S. Regulation of PCNA-protein interactions for genome stability. Nat. Rev. Mol. Cell Biol. 14, 269–282 (2013).

    Article  CAS  Google Scholar 

  170. Shenker, S. et al. Abstract 1337: Functional genomic characterization of the USP1 inhibitor KSQ-4279 reveals a distinct mechanism of action and resistance profile relative to other DDR targeting drugs. Cancer Res. 81, 1337–1337 (2021).

    Article  Google Scholar 

  171. ClinicalTrials.gov https://ClinicalTrials.gov/show/NCT05240898 (2022).

  172. Turner, N. C. et al. A synthetic lethal siRNA screen identifying genes mediating sensitivity to a PARP inhibitor. EMBO J. 27, 1368–1377 (2008).

    Article  CAS  Google Scholar 

  173. 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  Google Scholar 

  174. Neijenhuis, S., Bajrami, I., Miller, R., Lord, C. J. & Ashworth, A. Identification of miRNA modulators to PARP inhibitor response. DNA Repair 12, 394–402 (2013).

    Article  CAS  Google Scholar 

  175. Grundy, M. K., Buckanovich, R. J. & Bernstein, K. A. Regulation and pharmacological targeting of RAD51 in cancer. Nar. Cancer 2, zcaa024 (2020).

    Article  Google Scholar 

  176. Lamont, K. R. et al. Attenuating homologous recombination stimulates an AID-induced antileukemic effect. J. Exp. Med. 210, 1021–1033 (2013).

    Article  CAS  Google Scholar 

  177. Maclay, T. et al. CYT01B, a Novel RAD51 inhibitor, act synergistically with both targeted and chemotherapeutic anti-cancer agents. Blood https://doi.org/10.1182/blood-2018-99-119373 (2018).

    Article  Google Scholar 

  178. ClinicalTrials.gov https://ClinicalTrials.gov/show/NCT03997968 (2019).

  179. Vilar, E. & Gruber, S. B. Microsatellite instability in colorectal cancer-the stable evidence. Nat. Rev. Clin. Oncol. 7, 153–162 (2010).

    Article  CAS  Google Scholar 

  180. Cortes-Ciriano, I., Lee, S., Park, W. Y., Kim, T. M. & Park, P. J. A molecular portrait of microsatellite instability across multiple cancers. Nat. Commun. 8, 15180 (2017).

    Article  CAS  Google Scholar 

  181. Ratti, M., Lampis, A., Hahne, J. C., Passalacqua, R. & Valeri, N. Microsatellite instability in gastric cancer: molecular bases, clinical perspectives, and new treatment approaches. Cell Mol. Life Sci. 75, 4151–4162 (2018).

    Article  CAS  Google Scholar 

  182. Chan, E. M. et al. WRN helicase is a synthetic lethal target in microsatellite unstable cancers. Nature 568, 551–556 (2019).

    Article  CAS  Google Scholar 

  183. Behan, F. M. et al. Prioritization of cancer therapeutic targets using CRISPR-Cas9 screens. Nature 568, 511–516 (2019).

    Article  CAS  Google Scholar 

  184. van Wietmarschen, N. et al. Repeat expansions confer WRN dependence in microsatellite-unstable cancers. Nature 586, 292–298 (2020). This study shows that MSI rather than MMR deficiency alone renders cancer cells dependent on WRN.

    Article  Google Scholar 

  185. Tutt, A. N. et al. Exploiting the DNA repair defect in BRCA mutant cells in the design of new therapeutic strategies for cancer. Cold Spring Harb. Symp. Quant. Biol. 70, 139–148 (2005).

    Article  CAS  Google Scholar 

  186. Lindahl, T. & Barnes, D. E. Repair of endogenous DNA damage. Cold Spring Harb. Symp. Quant. Biol. 65, 127–133 (2000).

    Article  CAS  Google Scholar 

  187. Reislander, T., Groelly, F. J. & Tarsounas, M. DNA damage and cancer immunotherapy: a STING in the tale. Mol. Cell 80, 21–28 (2020).

    Article  CAS  Google Scholar 

  188. Ding, L. et al. PARP inhibition elicits STING-dependent antitumor immunity in Brca1-deficient ovarian cancer. Cell Rep. 25, 2972–2980 e2975 (2018).

    Article  CAS  Google Scholar 

  189. Pettitt, S. J. et al. Genome-wide and high-density CRISPR-Cas9 screens identify point mutations in PARP1 causing PARP inhibitor resistance. Nat. Commun. 9, 1849 (2018).

    Article  Google Scholar 

  190. Daniel, J. A. et al. Loss of ATM kinase activity leads to embryonic lethality in mice. J. Cell Biol. 198, 295–304 (2012).

    Article  CAS  Google Scholar 

  191. Yamamoto, K. et al. Kinase-dead ATM protein causes genomic instability and early embryonic lethality in mice. J. Cell Biol. 198, 305–313 (2012).

    Article  CAS  Google Scholar 

  192. Jiang, W. et al. Differential phosphorylation of DNA-PKcs regulates the interplay between end-processing and end-ligation during nonhomologous end-joining. Mol. Cell 58, 172–185 (2015).

    Article  CAS  Google Scholar 

  193. Koh, G., Degasperi, A., Zou, X., Momen, S. & Nik-Zainal, S. Mutational signatures: emerging concepts, caveats and clinical applications. Nat. Rev. Cancer 21, 619–637 (2021).

    Article  CAS  Google Scholar 

  194. Gourley, C. et al. Moving from Poly (ADP-Ribose) polymerase inhibition to targeting DNA repair and DNA damage response in cancer therapy. J. Clin. Oncol. 37, 2257–2269 (2019).

    Article  CAS  Google Scholar 

  195. Chabanon, R. M. et al. Targeting the DNA damage response in immuno-oncology: developments and opportunities. Nat. Rev. Cancer 21, 701–717 (2021). This Review covers the interplay between the DDR and the innate immune response.

    Article  CAS  Google Scholar 

  196. Yap, T. A. et al. First-in-human trial of the oral ataxia telangiectasia and RAD3-related (ATR) inhibitor BAY 1895344 in patients with advanced solid tumors. Cancer Discov. 11, 80–91 (2021).

    Article  CAS  Google Scholar 

  197. Konstantinopoulos, P. A. et al. Berzosertib plus gemcitabine versus gemcitabine alone in platinum-resistant high-grade serous ovarian cancer: a multicentre, open-label, randomised, phase 2 trial. Lancet Oncol. 21, 957–968 (2020).

    Article  CAS  Google Scholar 

  198. Hong, D. S. et al. Preclinical evaluation and phase Ib study of prexasertib, a CHK1 inhibitor, and samotolisib (LY3023414), a dual PI3K/mTOR inhibitor. Clin. Cancer Res. 27, 1864–1874 (2021).

    Article  CAS  Google Scholar 

  199. Moore, K. N. et al. Adavosertib with chemotherapy in patients with primary platinum-resistant ovarian, fallopian tube, or peritoneal cancer: an open-label, four-arm, phase II study. Clin. Cancer Res. 28, 36–44 (2022).

    Article  CAS  Google Scholar 

  200. Cuneo, K. C. et al. Dose escalation trial of the Wee1 inhibitor adavosertib (AZD1775) in combination with gemcitabine and radiation for patients with locally advanced pancreatic cancer. J. Clin. Oncol. 37, 2643–2650 (2019).

    Article  CAS  Google Scholar 

  201. Kristeleit, R. et al. Rucaparib versus standard-of-care chemotherapy in patients with relapsed ovarian cancer and a deleterious BRCA1 or BRCA2 mutation (ARIEL4): an international, open-label, randomised, phase 3 trial. Lancet Oncol. 23, 465–478 (2022).

    Article  CAS  Google Scholar 

  202. Bang, Y. J. et al. Olaparib in combination with paclitaxel in patients with advanced gastric cancer who have progressed following first-line therapy (GOLD): a double-blind, randomised, placebo-controlled, phase 3 trial. Lancet Oncol. 18, 1637–1651 (2017).

    Article  CAS  Google Scholar 

  203. Ramalingam, S. S. et al. Veliparib in combination with platinum-based chemotherapy for first-line treatment of advanced squamous cell lung cancer: a randomized, multicenter phase III study. J. Clin. Oncol. 39, 3633–3644 (2021).

    Article  CAS  Google Scholar 

  204. Govindan, R. et al. Veliparib plus carboplatin and paclitaxel versus investigator’s choice of standard chemotherapy in patients with advanced non-squamous non-small cell lung cancer. Clin. Lung Cancer 23, 214–225 (2022).

    Article  CAS  Google Scholar 

  205. Loibl, S. et al. Addition of the PARP inhibitor veliparib plus carboplatin or carboplatin alone to standard neoadjuvant chemotherapy in triple-negative breast cancer (BrighTNess): a randomised, phase 3 trial. Lancet Oncol. 19, 497–509 (2018).

    Article  CAS  Google Scholar 

  206. Dieras, V. et al. Veliparib with carboplatin and paclitaxel in BRCA-mutated advanced breast cancer (BROCADE3): a randomised, double-blind, placebo-controlled, phase 3 trial. Lancet Oncol. 21, 1269–1282 (2020).

    Article  CAS  Google Scholar 

  207. Coleman, R. L. et al. Veliparib with first-line chemotherapy and as maintenance therapy in ovarian cancer. N. Engl. J. Med. 381, 2403–2415 (2019).

    Article  CAS  Google Scholar 

  208. Ramalingam, S. S. et al. JASPER: phase 2 trial of first-line niraparib plus pembrolizumab in patients with advanced non-small cell lung cancer. Cancer 128, 65–74 (2022).

    Article  CAS  Google Scholar 

  209. Friedlander, M. et al. Pamiparib in combination with tislelizumab in patients with advanced solid tumours: results from the dose-escalation stage of a multicentre, open-label, phase 1a/b trial. Lancet Oncol. 20, 1306–1315 (2019).

    Article  CAS  Google Scholar 

  210. Sotiriou, S. K. et al. Mammalian RAD52 functions in break-induced replication repair of collapsed DNA replication forks. Mol. Cell 64, 1127–1134 (2016).

    Article  CAS  Google Scholar 

  211. Robson, M. et al. Olaparib for metastatic breast cancer in patients with a germline BRCA mutation. N. Engl. J. Med. 377, 523–533 (2017).

    Article  CAS  Google Scholar 

  212. Litton, J. K. et al. Talazoparib in patients with advanced breast cancer and a germline BRCA mutation. N. Engl. J. Med. 379, 753–763 (2018).

    Article  CAS  Google Scholar 

  213. Tutt, A. N. J. et al. Adjuvant olaparib for patients with BRCA1- or BRCA2-mutated breast cancer. N. Engl. J. Med. 384, 2394–2405 (2021).

    Article  CAS  Google Scholar 

  214. Moore, K. et al. Maintenance olaparib in patients with newly diagnosed advanced ovarian cancer. N. Engl. J. Med. 379, 2495–2505 (2018).

    Article  CAS  Google Scholar 

  215. Golan, T. et al. Maintenance olaparib for germline BRCA-mutated metastatic pancreatic cancer. N. Engl. J. Med. 381, 317–327 (2019).

    Article  CAS  Google Scholar 

  216. Abida, W. et al. Rucaparib in men with metastatic castration-resistant prostate cancer harboring a BRCA1 or BRCA2 gene alteration. J. Clin. Oncol. 38, 3763–3772 (2020).

    Article  CAS  Google Scholar 

  217. Moore, K. N. et al. Niraparib monotherapy for late-line treatment of ovarian cancer (QUADRA): a multicentre, open-label, single-arm, phase 2 trial. Lancet Oncol. 20, 636–648 (2019).

    Article  CAS  Google Scholar 

  218. Ray-Coquard, I. et al. Olaparib plus bevacizumab as first-line maintenance in ovarian cancer. N. Engl. J. Med. 381, 2416–2428 (2019).

    Article  CAS  Google Scholar 

  219. Hussain, M. et al. Survival with olaparib in metastatic castration-resistant prostate cancer. N. Engl. J. Med. 383, 2345–2357 (2020).

    Article  CAS  Google Scholar 

  220. Gonzalez-Martin, A. et al. Niraparib in patients with newly diagnosed advanced ovarian cancer. N. Engl. J. Med. 381, 2391–2402 (2019).

    Article  CAS  Google Scholar 

  221. Mirza, M. R. et al. Niraparib maintenance therapy in platinum-sensitive, recurrent ovarian cancer. N. Engl. J. Med. 375, 2154–2164 (2016).

    Article  CAS  Google Scholar 

  222. Ledermann, J. A. et al. Overall survival in patients with platinum-sensitive recurrent serous ovarian cancer receiving olaparib maintenance monotherapy: an updated analysis from a randomised, placebo-controlled, double-blind, phase 2 trial. Lancet Oncol. 17, 1579–1589 (2016).

    Article  CAS  Google Scholar 

  223. Ledermann, J. A. et al. Rucaparib for patients with platinum-sensitive, recurrent ovarian carcinoma (ARIEL3): post-progression outcomes and updated safety results from a randomised, placebo-controlled, phase 3 trial. Lancet Oncol. 21, 710–722 (2020).

    Article  CAS  Google Scholar 

  224. Lee, A. Fuzuloparib: first approval. Drugs 81, 1221–1226 (2021).

    Article  CAS  Google Scholar 

  225. Markham, A. Pamiparib: first approval. Drugs 81, 1343–1348 (2021).

    Article  CAS  Google Scholar 

  226. Zandarashvili, L. et al. Structural basis for allosteric PARP-1 retention on DNA breaks. Science https://doi.org/10.1126/science.aax6367 (2020).

    Article  Google Scholar 

  227. Hoppe, M. M., Sundar, R., Tan, D. S. P. & Jeyasekharan, A. D. Biomarkers for homologous recombination deficiency in cancer. J. Natl Cancer Inst. 110, 704–713 (2018).

    Article  Google Scholar 

  228. 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  Google Scholar 

  229. Castroviejo-Bermejo, M. et al. A RAD51 assay feasible in routine tumor samples calls PARP inhibitor response beyond BRCA mutation. EMBO Mol. Med. https://doi.org/10.15252/emmm.201809172 (2018).

    Article  Google Scholar 

  230. Cruz, C. et al. RAD51 foci as a functional biomarker of homologous recombination repair and PARP inhibitor resistance in germline BRCA-mutated breast cancer. Ann. Oncol. 29, 1203–1210 (2018). This study shows that a formation of RAD51 foci assay on diagnostic formalin-fixed, paraffin-embedded tumour samples can predict response to PARP inhibitors.

    Article  CAS  Google Scholar 

  231. Pettitt, S. J. et al. Clinical BRCA1/2 reversion analysis identifies hotspot mutations and predicted neoantigens associated with therapy resistance. Cancer Discov. 10, 1475–1488 (2020).

    Article  CAS  Google Scholar 

  232. Swisher, E. M. et al. Rucaparib in relapsed, platinum-sensitive high-grade ovarian carcinoma (ARIEL2 Part 1): an international, multicentre, open-label, phase 2 trial. Lancet Oncol. 18, 75–87 (2017).

    Article  CAS  Google Scholar 

  233. Ter Brugge, P. et al. Mechanisms of therapy resistance in patient-derived xenograft models of BRCA1-deficient breast cancer. J. Natl Cancer Inst. https://doi.org/10.1093/jnci/djw148 (2016).

    Article  Google Scholar 

  234. Jaspers, J. E. et al. Loss of 53BP1 causes PARP inhibitor resistance in Brca1-mutated mouse mammary tumors. Cancer Discov. 3, 68–81 (2013).

    Article  CAS  Google Scholar 

  235. Escribano-Diaz, C. et al. A cell cycle-dependent regulatory circuit composed of 53BP1-RIF1 and BRCA1-CtIP controls DNA repair pathway choice. Mol. Cell 49, 872–883 (2013).

    Article  CAS  Google Scholar 

  236. Zimmermann, M., Lottersberger, F., Buonomo, S. B., Sfeir, A. & de Lange, T. 53BP1 regulates DSB repair using Rif1 to control 5’ end resection. Science 339, 700–704 (2013).

    Article  CAS  Google Scholar 

  237. Di Virgilio, M. et al. Rif1 prevents resection of DNA breaks and promotes immunoglobulin class switching. Science 339, 711–715 (2013).

    Article  Google Scholar 

  238. Chapman, J. R. et al. RIF1 is essential for 53BP1-dependent nonhomologous end joining and suppression of DNA double-strand break resection. Mol. Cell 49, 858–871 (2013).

    Article  CAS  Google Scholar 

  239. Dev, H. et al. Shieldin complex promotes DNA end-joining and counters homologous recombination in BRCA1-null cells. Nat. Cell Biol. 20, 954–965 (2018).

    Article  CAS  Google Scholar 

  240. Gupta, R. et al. DNA repair network analysis reveals shieldin as a key regulator of NHEJ and PARP inhibitor sensitivity. Cell 173, 972–988 e923 (2018).

    Article  CAS  Google Scholar 

  241. Ghezraoui, H. et al. 53BP1 cooperation with the REV7-shieldin complex underpins DNA structure-specific NHEJ. Nature 560, 122–127 (2018).

    Article  CAS  Google Scholar 

  242. Noordermeer, S. M. et al. The shieldin complex mediates 53BP1-dependent DNA repair. Nature 560, 117–121 (2018).

    Article  CAS  Google Scholar 

  243. Gogola, E. et al. Selective loss of PARG restores PARylation and counteracts PARP inhibitor-mediated synthetic lethality. Cancer Cell 33, 1078–1093 (2018).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

The laboratories of the authors are supported by the University of Oxford and a Cancer Research UK Programme Award (DRCPGM\100001) to M.T. and a Cancer Research UK Career Development Fellowship (C29215/A20772) to A.N.B.

Author information

Authors and Affiliations

Authors

Contributions

All authors researched data for the article and contributed substantially to discussion of the content. F.J.G., A.N.B. and M.T. wrote the article and reviewed and/or edited the manuscript before submission.

Corresponding authors

Correspondence to Andrew N. Blackford or Madalena Tarsounas.

Ethics declarations

Competing interests

The authors declare no competing interests

Peer review

Peer review information

Nature Reviews Cancer thanks Nicola Curtin and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Related links

ClinicalTrials.gov: https://clinicaltrials.gov/

European Medicines Agency: https://www.ema.europa.eu/

Glossary

cGAS–STING axis

A pathway that detects cytosolic DNA and activates innate immune responses.

Chromothripsis

The process of chromosome shattering followed by incorrect repair.

Displacement loop

(D-loop). A DNA structure formed when a double helix is separated by invasion of a complementary single-stranded DNA end during homologous recombination.

Double Holliday junction

A four-way joint DNA molecule and homologous recombination intermediate that can form after second-end capture and ligation.

Micronuclei

Cytosolic structures containing a chromosome or chromosome fragment that are not incorporated in the nucleus during mitosis.

One-ended DSBs

Double-stranded DNA ends formed when a replication fork collapses, where there is no second DNA end available for ligation.

Second-end capture

The annealing step that pairs the second resected single-stranded DNA end from one side of a double-strand break with the joint molecule formed by invasion of a template DNA by the first resected DNA end during homologous recombination.

Shieldin complex

A protein complex consisting of SHLD1–SHLD2–SHLD3, which acts together with 53BP1–RIF1–REV7 to limit end resection at DNA double-strand breaks.

Two-ended DSBs

DNA double-strand breaks (DSBs) where both ends are available for ligation.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Groelly, F.J., Fawkes, M., Dagg, R.A. et al. Targeting DNA damage response pathways in cancer. Nat Rev Cancer 23, 78–94 (2023). https://doi.org/10.1038/s41568-022-00535-5

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41568-022-00535-5

  • Springer Nature Limited

This article is cited by

Navigation