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:

Exploiting replicative stress to treat cancer

Key Points

  • Tumour cells generally show enhanced replicative stress, a condition in which DNA polymerases lag behind unwinding DNA. These replication forks trigger a specific stress response.

  • Replicative stress can result from enhanced proliferative stimuli in tumour cells. It is accompanied by replication fork stalling, disassembly of the replication machinery and, ultimately, breakage of DNA.

  • Replicative stress can be further enhanced by chemotherapeutics, such as nucleoside analogues, topoisomerase inhibitors and DNA-modifying drugs.

  • Numerous signalling intermediates modulate the development of replicative stress and the cellular response to it. Many of these are established or potential drug targets to enhance replicative stress to eliminate tumour cells.

  • Anticancer therapy does not necessarily need to inhibit cell cycle progression. Rather, the general lack of growth control in cancer cells, which pushes them further through DNA replication, could be exploited to elicit a stress condition that promotes cancer cell death.

Abstract

DNA replication in cancer cells is accompanied by stalling and collapse of the replication fork and signalling in response to DNA damage and/or premature mitosis; these processes are collectively known as 'replicative stress'. Progress is being made to increase our understanding of the mechanisms that govern replicative stress, thus providing ample opportunities to enhance replicative stress for therapeutic purposes. Rather than trying to halt cell cycle progression, cancer therapeutics could aim to increase replicative stress by further loosening the checkpoints that remain available to cancer cells and ultimately inducing the catastrophic failure of proliferative machineries. In this Review, we outline current and future approaches to achieve this, emphasizing the combination of conventional chemotherapy with targeted approaches.

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: Generation of replicative stress and the resulting signalling cascades.
Figure 2: Generation and levels of replicative stress in tumour cells.
Figure 3: Timeline: a history of discoveries leading to drugs and drug candidates that increase replicative stress.
Figure 4: Targets to increase replicative stress.

Similar content being viewed by others

References

  1. Branzei, D. & Foiani, M. Maintaining genome stability at the replication fork. Nature Rev. Mol. Cell Biol. 11, 208–219 (2010).

    CAS  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Sabatinos, S. A. Recovering a stalled replication fork. Nature Educ. 3, 31 (2010).

    Google Scholar 

  4. Zeman, M. K. & Cimprich, K. A. Causes and consequences of replication stress. Nature Cell Biol. 16, 2–9 (2014).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  6. Gorgoulis, V. G. et al. Activation of the DNA damage checkpoint and genomic instability in human precancerous lesions. Nature 434, 907–913 (2005).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  8. Bartkova, J. et al. Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444, 633–637 (2006). References 7 and 8 show that replicative stress is specifically found in tumour cells.

    CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  10. Hanahan, D. & Weinberg, R. A. The hallmarks of cancer. Cell 100, 57–70 (2000).

    CAS  PubMed  Google Scholar 

  11. Gorgoulis, V. G. & Halazonetis, T. D. Oncogene-induced senescence: the bright and dark side of the response. Curr. Opin. Cell Biol. 22, 816–827 (2010).

    CAS  PubMed  Google Scholar 

  12. Negrini, S., Gorgoulis, V. G. & Halazonetis, T. D. Genomic instability — an evolving hallmark of cancer. Nature Rev. Mol. Cell Biol. 11, 220–228 (2010).

    CAS  Google Scholar 

  13. Bester, A. C. et al. Nucleotide deficiency promotes genomic instability in early stages of cancer development. Cell 145, 435–446 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Tort, F. et al. Retinoblastoma pathway defects show differential ability to activate the constitutive DNA damage response in human tumorigenesis. Cancer Res. 66, 10258–10263 (2006).

    CAS  PubMed  Google Scholar 

  15. Vafa, O. et al. c-Myc can induce DNA damage, increase reactive oxygen species, and mitigate p53 function: a mechanism for oncogene-induced genetic instability. Mol. Cell 9, 1031–1044 (2002).

    CAS  PubMed  Google Scholar 

  16. Sabharwal, S. S. & Schumacker, P. T. Mitochondrial ROS in cancer: initiators, amplifiers or an Achilles' heel? Nature Rev. Cancer 14, 709–721 (2014).

    CAS  Google Scholar 

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

    CAS  Google Scholar 

  18. Helleday, T. Cancer phenotypic lethality, exemplified by the non-essential MTH1 enzyme being required for cancer survival. Ann. Oncol. 25, 1253–1255 (2014).

    CAS  PubMed  Google Scholar 

  19. Huber, K. V. et al. Stereospecific targeting of MTH1 by (S)-crizotinib as an anticancer strategy. Nature 508, 222–227 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Gad, H. et al. MTH1 inhibition eradicates cancer by preventing sanitation of the dNTP pool. Nature 508, 215–221 (2014).

    CAS  PubMed  Google Scholar 

  21. Markkanen, E., Castrec, B., Villani, G. & Hubscher, U. A switch between DNA polymerases δ and λ promotes error-free bypass of 8-oxo-G lesions. Proc. Natl Acad. Sci. USA 109, 20401–20406 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Sung, J. S. & Demple, B. Roles of base excision repair subpathways in correcting oxidized abasic sites in DNA. FEBS J. 273, 1620–1629 (2006).

    CAS  PubMed  Google Scholar 

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

    CAS  Google Scholar 

  24. Vogelstein, B. & Kinzler, K. W. Cancer genes and the pathways they control. Nature Med. 10, 789–799 (2004).

    CAS  PubMed  Google Scholar 

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

    CAS  Google Scholar 

  26. Pathania, S. et al. BRCA1 is required for postreplication repair after UV-induced DNA damage. Mol. Cell 44, 235–251 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Nikkila, J. et al. Heterozygous mutations in PALB2 cause DNA replication and damage response defects. Nature Commun. 4, 2578 (2013).

    Google Scholar 

  28. Tanaka, H. et al. A ribonucleotide reductase gene involved in a p53-dependent cell-cycle checkpoint for DNA damage. Nature 404, 42–49 (2000).

    CAS  PubMed  Google Scholar 

  29. Nakano, K., Balint, E., Ashcroft, M. & Vousden, K. H. A ribonucleotide reductase gene is a transcriptional target of p53 and p73. Oncogene 19, 4283–4289 (2000).

    CAS  PubMed  Google Scholar 

  30. Borlado, L. R. & Mendez, J. CDC6: from DNA replication to cell cycle checkpoints and oncogenesis. Carcinogenesis 29, 237–243 (2008).

    CAS  PubMed  Google Scholar 

  31. Liontos, M. et al. Deregulated overexpression of hCdt1 and hCdc6 promotes malignant behavior. Cancer Res. 67, 10899–10909 (2007).

    CAS  PubMed  Google Scholar 

  32. Hoffmann, J. S. & Cazaux, C. Aberrant expression of alternative DNA polymerases: a source of mutator phenotype as well as replicative stress in cancer. Semin. Cancer Biol. 20, 312–319 (2010).

    CAS  PubMed  Google Scholar 

  33. Santarius, T., Shipley, J., Brewer, D., Stratton, M. R. & Cooper, C. S. A census of amplified and overexpressed human cancer genes. Nature Rev. Cancer 10, 59–64 (2010).

    CAS  Google Scholar 

  34. Burhans, W. C. & Weinberger, M. DNA replication stress, genome instability and aging. Nucleic Acids Res. 35, 7545–7556 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Fu, D., Calvo, J. A. & Samson, L. D. Balancing repair and tolerance of DNA damage caused by alkylating agents. Nature Rev. Cancer 12, 104–120 (2012).

    CAS  Google Scholar 

  36. Wang, D. & Lippard, S. J. Cellular processing of platinum anticancer drugs. Nature Rev. Drug Discov. 4, 307–320 (2005).

    CAS  Google Scholar 

  37. Henry-Mowatt, J. et al. XRCC3 and Rad51 modulate replication fork progression on damaged vertebrate chromosomes. Mol. Cell 11, 1109–1117 (2003).

    CAS  PubMed  Google Scholar 

  38. Sale, J. E., Lehmann, A. R. & Woodgate, R. Y-family DNA polymerases and their role in tolerance of cellular DNA damage. Nature Rev. Mol. Cell Biol. 13, 141–152 (2012).

    CAS  Google Scholar 

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

    CAS  Google Scholar 

  40. Sedgwick, B. Repairing DNA-methylation damage. Nature Rev. Mol. Cell Biol. 5, 148–157 (2004).

    CAS  Google Scholar 

  41. Kopper, F. et al. Damage-induced DNA replication stalling relies on MAPK-activated protein kinase 2 activity. Proc. Natl Acad. Sci. USA 110, 16856–16861 (2013). This study shows that MK2 is a mediator of replicative stress in nucleoside analogue treatment.

    PubMed  PubMed Central  Google Scholar 

  42. Merrick, C. J., Jackson, D. & Diffley, J. F. Visualization of altered replication dynamics after DNA damage in human cells. J. Biol. Chem. 279, 20067–20075 (2004).

    CAS  PubMed  Google Scholar 

  43. Ewald, B., Sampath, D. & Plunkett, W. Nucleoside analogs: molecular mechanisms signaling cell death. Oncogene 27, 6522–6537 (2008).

    CAS  PubMed  Google Scholar 

  44. Orta, M. L. et al. 5-aza-2′-deoxycytidine causes replication lesions that require Fanconi anemia-dependent homologous recombination for repair. Nucleic Acids Res. 41, 5827–5836 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Longley, D. B., Harkin, D. P. & Johnston, P. G. 5-fluorouracil: mechanisms of action and clinical strategies. Nature Rev. Cancer 3, 330–338 (2003).

    CAS  Google Scholar 

  46. Huang, P. & Plunkett, W. Action of 9-β-D-arabinofuranosyl-2-fluoroadenine on RNA metabolism. Mol. Pharmacol. 39, 449–455 (1991).

    CAS  PubMed  Google Scholar 

  47. Donati, G., Peddigari, S., Mercer, C. A. & Thomas, G. 5S ribosomal RNA is an essential component of a nascent ribosomal precursor complex that regulates the Hdm2–p53 checkpoint. Cell Rep. 4, 87–98 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Pommier, Y. Drugging topoisomerases: lessons and challenges. ACS Chem. Biol. 8, 82–95 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Regairaz, M. et al. Mus81-mediated DNA cleavage resolves replication forks stalled by topoisomerase I–DNA complexes. J. Cell Biol. 195, 739–749 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Seiler, J. A., Conti, C., Syed, A., Aladjem, M. I. & Pommier, Y. The intra-S-phase checkpoint affects both DNA replication initiation and elongation: single-cell and -DNA fiber analyses. Mol. Cell. Biol. 27, 5806–5818 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Ray Chaudhuri, A. et al. Topoisomerase I poisoning results in PARP-mediated replication fork reversal. Nature Struct. Mol. Biol. 19, 417–423 (2012).

    CAS  Google Scholar 

  52. Berti, M. et al. Human RECQ1 promotes restart of replication forks reversed by DNA topoisomerase I inhibition. Nature Struct. Mol. Biol. 20, 347–354 (2013).

    CAS  Google Scholar 

  53. Rodriguez, R. & Meuth, M. Chk1 and p21 cooperate to prevent apoptosis during DNA replication fork stress. Mol. Biol. Cell 17, 402–412 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Loegering, D. et al. Rad9 protects cells from topoisomerase poison-induced cell death. J. Biol. Chem. 279, 18641–18647 (2004).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  56. Ledermann, J. et al. Olaparib maintenance therapy in platinum-sensitive relapsed ovarian cancer. N. Engl. J. Med. 366, 1382–1392 (2012).

    CAS  PubMed  Google Scholar 

  57. Altmeyer, M. et al. The chromatin scaffold protein SAFB1 renders chromatin permissive for DNA damage signaling. Mol. Cell 52, 206–220 (2013).

    CAS  PubMed  Google Scholar 

  58. Min, W. et al. Poly(ADP-ribose) binding to Chk1 at stalled replication forks is required for S-phase checkpoint activation. Nature Commun. 4, 2993 (2013).

    Google Scholar 

  59. Murai, J. et al. Rationale for poly(ADP-ribose) polymerase (PARP) inhibitors in combination therapy with camptothecins or temozolomide based on PARP trapping versus catalytic inhibition. J. Pharmacol. Exp. Ther. 349, 408–416 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Zhou, B. B. & Bartek, J. Targeting the checkpoint kinases: chemosensitization versus chemoprotection. Nature Rev. Cancer 4, 216–225 (2004).

    CAS  Google Scholar 

  61. Toledo, L. I., Murga, M. & Fernandez-Capetillo, O. Targeting ATR and Chk1 kinases for cancer treatment: a new model for new (and old) drugs. Mol. Oncol. 5, 368–373 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Syljuasen, R. G. et al. Inhibition of human CHK1 causes increased initiation of DNA replication, phosphorylation of ATR targets, and DNA breakage. Mol. Cell. Biol. 25, 3553–3562 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Petermann, E., Woodcock, M. & Helleday, T. Chk1 promotes replication fork progression by controlling replication initiation. Proc. Natl Acad. Sci. USA 107, 16090–16095 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Lee, J., Kumagai, A. & Dunphy, W. G. Claspin, a Chk1-regulatory protein, monitors DNA replication on chromatin independently of RPA, ATR, and Rad17. Mol. Cell 11, 329–340 (2003).

    CAS  PubMed  Google Scholar 

  65. Kumagai, A. & Dunphy, W. G. Repeated phosphopeptide motifs in claspin mediate the regulated binding of Chk1. Nature Cell Biol. 5, 161–165 (2003).

    CAS  PubMed  Google Scholar 

  66. Ma, C. X. et al. A phase II study of UCN-01 in combination with irinotecan in patients with metastatic triple negative breast cancer. Breast Cancer Res. Treat. 137, 483–492 (2013).

    CAS  PubMed  Google Scholar 

  67. Li, T. et al. A phase II study of cell cycle inhibitor UCN-01 in patients with metastatic melanoma: a California cancer consortium trial. Invest. New Drugs 30, 741–748 (2012).

    PubMed  Google Scholar 

  68. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Sorensen, C. S. & Syljuasen, R. G. Safeguarding genome integrity: the checkpoint kinases ATR, CHK1 and WEE1 restrain CDK activity during normal DNA replication. Nucleic Acids Res. 40, 477–486 (2012).

    CAS  PubMed  Google Scholar 

  70. Couch, F. B. et al. ATR phosphorylates SMARCAL1 to prevent replication fork collapse. Genes Dev. 27, 1610–1623 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. 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).

    CAS  Google Scholar 

  72. Aarts, M. et al. Forced mitotic entry of S-phase cells as a therapeutic strategy induced by inhibition of WEE1. Cancer Discov. 2, 524–539 (2012).

    CAS  PubMed  Google Scholar 

  73. Beck, H. et al. Cyclin-dependent kinase suppression by WEE1 kinase protects the genome through control of replication initiation and nucleotide consumption. Mol. Cell. Biol. 32, 4226–4236 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Do, K., Doroshow, J. H. & Kummar, S. Wee1 kinase as a target for cancer therapy. Cell Cycle 12, 3159–3164 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Mollapour, M. et al. Swe1Wee1-dependent tyrosine phosphorylation of Hsp90 regulates distinct facets of chaperone function. Mol. Cell 37, 333–343 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  76. Mollapour, M., Tsutsumi, S. & Neckers, L. Hsp90 phosphorylation, Wee1 and the cell cycle. Cell Cycle 9, 2310–2316 (2010).

    CAS  PubMed  Google Scholar 

  77. Soucy, T. A. et al. An inhibitor of NEDD8-activating enzyme as a new approach to treat cancer. Nature 458, 732–736 (2009).

    CAS  PubMed  Google Scholar 

  78. Milhollen, M. A. et al. Inhibition of NEDD8-activating enzyme induces rereplication and apoptosis in human tumor cells consistent with deregulating CDT1 turnover. Cancer Res. 71, 3042–3051 (2011).

    CAS  PubMed  Google Scholar 

  79. Lin, J. J., Milhollen, M. A., Smith, P. G., Narayanan, U. & Dutta, A. NEDD8-targeting drug MLN4924 elicits DNA rereplication by stabilizing Cdt1 in S phase, triggering checkpoint activation, apoptosis, and senescence in cancer cells. Cancer Res. 70, 10310–10320 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  80. Tanaka, T., Nakatani, T. & Kamitani, T. Negative regulation of NEDD8 conjugation pathway by novel molecules and agents for anticancer therapy. Curr. Pharm. Des. 19, 4131–4139 (2013).

    CAS  PubMed  Google Scholar 

  81. Nawrocki, S. T., Griffin, P., Kelly, K. R. & Carew, J. S. MLN4924: a novel first-in-class inhibitor of NEDD8-activating enzyme for cancer therapy. Expert Opin. Investig. Drugs 21, 1563–1573 (2012).

    CAS  PubMed  Google Scholar 

  82. Milhollen, M. A. et al. Treatment-emergent mutations in NAEβ confer resistance to the NEDD8-activating enzyme inhibitor MLN4924. Cancer Cell 21, 388–401 (2012).

    CAS  PubMed  Google Scholar 

  83. Toth, J. I., Yang, L., Dahl, R. & Petroski, M. D. A gatekeeper residue for NEDD8-activating enzyme inhibition by MLN4924. Cell Rep. 1, 309–316 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  84. Gray, D. et al. Maternal embryonic leucine zipper kinase/murine protein serine-threonine kinase 38 is a promising therapeutic target for multiple cancers. Cancer Res. 65, 9751–9761 (2005).

    CAS  PubMed  Google Scholar 

  85. Hebbard, L. W. et al. Maternal embryonic leucine zipper kinase is upregulated and required in mammary tumor-initiating cells in vivo. Cancer Res. 70, 8863–8873 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Chung, S. & Nakamura, Y. MELK inhibitor, novel molecular targeted therapeutics for human cancer stem cells. Cell Cycle 12, 1655–1656 (2013).

    CAS  PubMed  Google Scholar 

  87. Kig, C. et al. Maternal embryonic leucine zipper kinase (MELK) reduces replication stress in glioblastoma cells. J. Biol. Chem. 288, 24200–24212 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Chung, S. et al. Development of an orally-administrative MELK-targeting inhibitor that suppresses the growth of various types of human cancer. Oncotarget 3, 1629–1640 (2012).

    PubMed  PubMed Central  Google Scholar 

  89. Cho, Y. S., Kang, Y., Kim, K., Cha, Y. J. & Cho, H. S. The crystal structure of MPK38 in complex with OTSSP167, an orally administrative MELK selective inhibitor. Biochem. Biophys. Res. Commun. 447, 7–11 (2014).

    CAS  PubMed  Google Scholar 

  90. Sheu, Y. J. & Stillman, B. The Dbf4–Cdc7 kinase promotes S phase by alleviating an inhibitory activity in Mcm4. Nature 463, 113–117 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Chuang, L. C. et al. Phosphorylation of Mcm2 by Cdc7 promotes pre-replication complex assembly during cell-cycle re-entry. Mol. Cell 35, 206–216 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Montagnoli, A. et al. A Cdc7 kinase inhibitor restricts initiation of DNA replication and has antitumor activity. Nature Chem. Biol. 4, 357–365 (2008).

    CAS  Google Scholar 

  93. Rodriguez-Acebes, S. et al. Targeting DNA replication before it starts: Cdc7 as a therapeutic target in p53-mutant breast cancers. Am. J. Pathol. 177, 2034–2045 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Montagnoli, A., Moll, J. & Colotta, F. Targeting cell division cycle 7 kinase: a new approach for cancer therapy. Clin. Cancer Res. 16, 4503–4508 (2010).

    CAS  PubMed  Google Scholar 

  95. Montagnoli, A. et al. Cdc7 inhibition reveals a p53-dependent replication checkpoint that is defective in cancer cells. Cancer Res. 64, 7110–7116 (2004).

    CAS  PubMed  Google Scholar 

  96. Yamada, M. et al. ATR–Chk1–APC/CCdh1-dependent stabilization of Cdc7–ASK (Dbf4) kinase is required for DNA lesion bypass under replication stress. Genes Dev. 27, 2459–2472 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  97. Yamada, M., Masai, H. & Bartek, J. Regulation and roles of Cdc7 kinase under replication stress. Cell Cycle 13, 1859–1866 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Filippakopoulos, P. & Knapp, S. Targeting bromodomains: epigenetic readers of lysine acetylation. Nature Rev. Drug Discov. 13, 337–356 (2014).

    CAS  Google Scholar 

  99. Filippakopoulos, P. et al. Selective inhibition of BET bromodomains. Nature 468, 1067–1073 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Da Costa, D. et al. BET inhibition as a single or combined therapeutic approach in primary paediatric B-precursor acute lymphoblastic leukaemia. Blood Cancer J. 3, e126 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  101. Bolden, J. E., Peart, M. J. & Johnstone, R. W. Anticancer activities of histone deacetylase inhibitors. Nature Rev. Drug Discov. 5, 769–784 (2006).

    CAS  Google Scholar 

  102. Conti, C. et al. Inhibition of histone deacetylase in cancer cells slows down replication forks, activates dormant origins, and induces DNA damage. Cancer Res. 70, 4470–4480 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  103. Ewald, B., Sampath, D. & Plunkett, W. H2AX phosphorylation marks gemcitabine-induced stalled replication forks and their collapse upon S-phase checkpoint abrogation. Mol. Cancer Ther. 6, 1239–1248 (2007).

    CAS  PubMed  Google Scholar 

  104. Prevo, R. et al. The novel ATR inhibitor VE-821 increases sensitivity of pancreatic cancer cells to radiation and chemotherapy. Cancer Biol. Ther. 13, 1072–1081 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  105. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  106. 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).

    CAS  PubMed  Google Scholar 

  107. Garcia, K. et al. Nedd8-activating enzyme inhibitor MLN4924 provides synergy with mitomycin C through interactions with ATR, BRCA1/BRCA2 and chromatin dynamics pathways. Mol. Cancer Ther. 13, 1625–1635 (2014).

    CAS  PubMed  Google Scholar 

  108. Sausville, E. et al. Phase I dose-escalation study of AZD7762, a checkpoint kinase inhibitor, in combination with gemcitabine in US patients with advanced solid tumors. Cancer Chemother. Pharmacol. 73, 539–549 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Seto, T. et al. Phase I, dose-escalation study of AZD7762 alone and in combination with gemcitabine in Japanese patients with advanced solid tumours. Cancer Chemother. Pharmacol. 72, 619–627 (2013).

    CAS  PubMed  Google Scholar 

  110. Perez, R. P. et al. Modulation of cell cycle progression in human tumors: a pharmacokinetic and tumor molecular pharmacodynamic study of cisplatin plus the Chk1 inhibitor UCN-01 (NSC 638850). Clin. Cancer Res. 12, 7079–7085 (2006).

    CAS  PubMed  Google Scholar 

  111. Yazlovitskaya, E. M. & Persons, D. L. Inhibition of cisplatin-induced ATR activity and enhanced sensitivity to cisplatin. Anticancer Res. 23, 2275–2279 (2003).

    CAS  PubMed  Google Scholar 

  112. Sangster-Guity, N., Conrad, B. H., Papadopoulos, N. & Bunz, F. ATR mediates cisplatin resistance in a p53 genotype-specific manner. Oncogene 30, 2526–2533 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Kee, Y. et al. Inhibition of the Nedd8 system sensitizes cells to DNA interstrand cross-linking agents. Mol. Cancer Res. 10, 369–377 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  114. Matei, D. et al. Epigenetic resensitization to platinum in ovarian cancer. Cancer Res. 72, 2197–2205 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  115. Reinhardt, H. C. & Yaffe, M. B. Kinases that control the cell cycle in response to DNA damage: Chk1, Chk2, and MK2. Curr. Opin. Cell Biol. 21, 245–255 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Morandell, S. et al. A reversible gene-targeting strategy identifies synthetic lethal interactions between MK2 and p53 in the DNA damage response in vivo. Cell Rep. 5, 868–877 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  117. Reinhardt, H. C. et al. DNA damage activates a spatially distinct late cytoplasmic cell-cycle checkpoint network controlled by MK2-mediated RNA stabilization. Mol. Cell 40, 34–49 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  118. Reinhardt, H. C., Aslanian, A. S., Lees, J. A. & Yaffe, M. B. p53-deficient cells rely on ATM- and ATR-mediated checkpoint signaling through the p38MAPK/MK2 pathway for survival after DNA damage. Cancer Cell 11, 175–189 (2007). References 116–118 show that MK2 is an inhibitor of cell death following platinum treatment.

    CAS  PubMed  PubMed Central  Google Scholar 

  119. Kopper, F., Binkowski, A. M., Bierwirth, C. & Dobbelstein, M. The MAPK-activated protein kinase 2 mediates gemcitabine sensitivity in pancreatic cancer cells. Cell Cycle 13, 884–889 (2014).

    PubMed  PubMed Central  Google Scholar 

  120. Chugh, R. et al. A preclinical evaluation of Minnelide as a therapeutic agent against pancreatic cancer. Sci. Transl Med. 4, 156ra139 (2012). This study shows that a MKP1-targeting drug is suitable for successful treatment in a cancer model.

    PubMed  PubMed Central  Google Scholar 

  121. Arlander, S. J. et al. Hsp90 inhibition depletes Chk1 and sensitizes tumor cells to replication stress. J. Biol. Chem. 278, 52572–52577 (2003).

    CAS  PubMed  Google Scholar 

  122. Ha, K. et al. Hsp90 inhibitor-mediated disruption of chaperone association of ATR with Hsp90 sensitizes cancer cells to DNA damage. Mol. Cancer Ther. 10, 1194–1206 (2011).

    CAS  PubMed  Google Scholar 

  123. Dobbelstein, M. & Moll, U. Targeting tumour-supportive cellular machineries in anticancer drug development. Nature Rev. Drug Discov. 13, 179–196 (2014).

    CAS  Google Scholar 

  124. Davies, K. D. et al. Chk1 inhibition and Wee1 inhibition combine synergistically to impede cellular proliferation. Cancer Biol. Ther. 12, 788–796 (2011).

    CAS  PubMed  Google Scholar 

  125. Russell, M. R. et al. Combination therapy targeting the Chk1 and Wee1 kinases shows therapeutic efficacy in neuroblastoma. Cancer Res. 73, 776–784 (2013).

    CAS  PubMed  Google Scholar 

  126. Guertin, A. D. et al. Unique functions of CHK1 and WEE1 underlie synergistic anti-tumor activity upon pharmacologic inhibition. Cancer Cell Int. 12, 45 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  127. Carrassa, L. et al. Combined inhibition of Chk1 and Wee1: in vitro synergistic effect translates to tumor growth inhibition in vivo. Cell Cycle 11, 2507–2517 (2012).

    CAS  PubMed  Google Scholar 

  128. Chaudhuri, L. et al. CHK1 and WEE1 inhibition combine synergistically to enhance therapeutic efficacy in acute myeloid leukemia ex vivo. Haematologica 99, 688–696 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  129. Sorensen, C. S. et al. The cell-cycle checkpoint kinase Chk1 is required for mammalian homologous recombination repair. Nature Cell Biol. 7, 195–201 (2005). References 63 and 129 show that inhibition of the signalling kinase CHK1 enhances replicative stress.

    CAS  PubMed  Google Scholar 

  130. Hu, B. et al. Fhit and CHK1 have opposing effects on homologous recombination repair. Cancer Res. 65, 8613–8616 (2005).

    CAS  PubMed  Google Scholar 

  131. Anderson, V. E. et al. CCT241533 is a potent and selective inhibitor of CHK2 that potentiates the cytotoxicity of PARP inhibitors. Cancer Res. 71, 463–472 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  132. 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).

    CAS  Google Scholar 

  133. Meek, D. W. Tumour suppression by p53: a role for the DNA damage response? Nature Rev. Cancer 9, 714–723 (2009).

    CAS  PubMed  Google Scholar 

  134. Khoo, K. H., Verma, C. S. & Lane, D. P. Drugging the p53 pathway: understanding the route to clinical efficacy. Nature Rev. Drug Discov. 13, 217–236 (2014).

    CAS  Google Scholar 

  135. Polager, S. & Ginsberg, D. p53 and E2f: partners in life and death. Nature Rev. Cancer 9, 738–748 (2009).

    CAS  Google Scholar 

  136. Vassilev, L. T. et al. In vivo activation of the p53 pathway by small-molecule antagonists of MDM2. Science 303, 844–848 (2004).

    CAS  PubMed  Google Scholar 

  137. Kranz, D. & Dobbelstein, M. Nongenotoxic p53 activation protects cells against S-phase-specific chemotherapy. Cancer Res. 66, 10274–10280 (2006).

    CAS  PubMed  Google Scholar 

  138. Kranz, D., Dohmesen, C. & Dobbelstein, M. BRCA1 and Tip60 determine the cellular response to ultraviolet irradiation through distinct pathways. J. Cell Biol. 182, 197–213 (2008). References 137 and 138 show that p53 protects cells from replicative stress.

    CAS  PubMed  PubMed Central  Google Scholar 

  139. Carvajal, D. et al. Activation of p53 by MDM2 antagonists can protect proliferating cells from mitotic inhibitors. Cancer Res. 65, 1918–1924 (2005).

    CAS  PubMed  Google Scholar 

  140. Yuan, Z. M. et al. p73 is regulated by tyrosine kinase c-Abl in the apoptotic response to DNA damage. Nature 399, 814–817 (1999).

    CAS  PubMed  Google Scholar 

  141. Agami, R., Blandino, G., Oren, M. & Shaul, Y. Interaction of c-Abl and p73α and their collaboration to induce apoptosis. Nature 399, 809–813 (1999).

    CAS  PubMed  Google Scholar 

  142. Gong, J. G. et al. The tyrosine kinase c-Abl regulates p73 in apoptotic response to cisplatin-induced DNA damage. Nature 399, 806–809 (1999).

    CAS  PubMed  Google Scholar 

  143. Beyer, U., Moll-Rocek, J., Moll, U. M. & Dobbelstein, M. Endogenous retrovirus drives hitherto unknown proapoptotic p63 isoforms in the male germ line of humans and great apes. Proc. Natl Acad. Sci. USA 108, 3624–3629 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  144. Lapenna, S. & Giordano, A. Cell cycle kinases as therapeutic targets for cancer. Nature Rev. Drug Discov. 8, 547–566 (2009).

    CAS  Google Scholar 

  145. [No authors listed.] CDK inhibitors speed ahead. Nature Rev. Drug Discov. 13, 323 (2014).

  146. Matranga, C. B. & Shapiro, G. I. Selective sensitization of transformed cells to flavopiridol-induced apoptosis following recruitment to S-phase. Cancer Res. 62, 1707–1717 (2002).

    CAS  PubMed  Google Scholar 

  147. Ali, S., El-Rayes, B. F., Aranha, O., Sarkar, F. H. & Philip, P. A. Sequence dependent potentiation of gemcitabine by flavopiridol in human breast cancer cells. Breast Cancer Res. Treat. 90, 25–31 (2005).

    CAS  PubMed  Google Scholar 

  148. Cayrol, C. & Ducommun, B. Interaction with cyclin-dependent kinases and PCNA modulates proteasome-dependent degradation of p21. Oncogene 17, 2437–2444 (1998).

    CAS  PubMed  Google Scholar 

  149. Kossatz, U. et al. Skp2-dependent degradation of p27kip1 is essential for cell cycle progression. Genes Dev. 18, 2602–2607 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  150. Nickeleit, I. et al. Argyrin a reveals a critical role for the tumor suppressor protein p27kip1 in mediating antitumor activities in response to proteasome inhibition. Cancer Cell 14, 23–35 (2008).

    CAS  PubMed  Google Scholar 

  151. Fahy, B. N., Schlieman, M. G., Virudachalam, S. & Bold, R. J. Schedule-dependent molecular effects of the proteasome inhibitor bortezomib and gemcitabine in pancreatic cancer. J. Surg. Res. 113, 88–95 (2003).

    CAS  PubMed  Google Scholar 

  152. Pattabiraman, D. R. & Weinberg, R. A. Tackling the cancer stem cells — what challenges do they pose? Nature Rev. Drug Discov. 13, 497–512 (2014).

    CAS  Google Scholar 

  153. Wang, W. J. et al. MYC regulation of CHK1 and CHK2 promotes radioresistance in a stem cell-like population of nasopharyngeal carcinoma cells. Cancer Res. 73, 1219–1231 (2013).

    CAS  PubMed  Google Scholar 

  154. Hoglund, A. et al. Therapeutic implications for the induced levels of Chk1 in Myc-expressing cancer cells. Clin. Cancer Res. 17, 7067–7079 (2011).

    PubMed  Google Scholar 

  155. Ragland, R. L. et al. RNF4 and PLK1 are required for replication fork collapse in ATR-deficient cells. Genes Dev. 27, 2259–2273 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  156. Helleday, T., Petermann, E., Lundin, C., Hodgson, B. & Sharma, R. A. DNA repair pathways as targets for cancer therapy. Nature Rev. Cancer 8, 193–204 (2008). This study highlights DNA repair pathways as a target for cancer therapy.

    CAS  Google Scholar 

  157. Aapro, M. S., Martin, C. & Hatty, S. Gemcitabine — a safety review. Anticancer Drugs 9, 191–201 (1998).

    CAS  PubMed  Google Scholar 

  158. Chabner, B. A. & Roberts, T. G. Jr. Chemotherapy and the war on cancer. Nature Rev. Cancer 5, 65–72 (2005).

    CAS  Google Scholar 

  159. Corey, S. J. et al. Myelodysplastic syndromes: the complexity of stem-cell diseases. Nature Rev. Cancer 7, 118–129 (2007).

    CAS  Google Scholar 

  160. Costantino, L. et al. Break-induced replication repair of damaged forks induces genomic duplications in human cells. Science 343, 88–91 (2014).

    CAS  PubMed  Google Scholar 

  161. van Leeuwen, I. M., Rao, B., Sachweh, M. C. & Lain, S. An evaluation of small-molecule p53 activators as chemoprotectants ameliorating adverse effects of anticancer drugs in normal cells. Cell Cycle 11, 1851–1861 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  162. Rao, B., Lain, S. & Thompson, A. M. p53-based cyclotherapy: exploiting the 'guardian of the genome' to protect normal cells from cytotoxic therapy. Br. J. Cancer 109, 2954–2958 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  163. Blagosklonny, M. V. Wt p53 impairs response to chemotherapy: make lemonade to spare normal cells. Oncotarget 3, 601–607 (2012).

    PubMed  PubMed Central  Google Scholar 

  164. Ray-Coquard, I. et al. Effect of the MDM2 antagonist RG7112 on the P53 pathway in patients with MDM2-amplified, well-differentiated or dedifferentiated liposarcoma: an exploratory proof-of-mechanism study. Lancet Oncol. 13, 1133–1140 (2012).

    CAS  PubMed  Google Scholar 

  165. Falck, J., Mailand, N., Syljuasen, R. G., Bartek, J. & Lukas, J. The ATM–Chk2–Cdc25A checkpoint pathway guards against radioresistant DNA synthesis. Nature 410, 842–847 (2001).

    CAS  PubMed  Google Scholar 

  166. Falck, J., Petrini, J. H., Williams, B. R., Lukas, J. & Bartek, J. The DNA damage-dependent intra-S phase checkpoint is regulated by parallel pathways. Nature Genet. 30, 290–294 (2002).

    PubMed  Google Scholar 

  167. Sorensen, C. S. et al. Chk1 regulates the S phase checkpoint by coupling the physiological turnover and ionizing radiation-induced accelerated proteolysis of Cdc25A. Cancer Cell 3, 247–258 (2003).

    CAS  PubMed  Google Scholar 

  168. Koniaras, K., Cuddihy, A. R., Christopoulos, H., Hogg, A. & O'Connell, M. J. Inhibition of Chk1-dependent G2 DNA damage checkpoint radiosensitizes p53 mutant human cells. Oncogene 20, 7453–7463 (2001).

    CAS  PubMed  Google Scholar 

  169. Hu, B. et al. The radioresistance to killing of A1–5 cells derives from activation of the Chk1 pathway. J. Biol. Chem. 276, 17693–17698 (2001).

    CAS  PubMed  Google Scholar 

  170. Fernet, M., Megnin-Chanet, F., Hall, J. & Favaudon, V. Control of the G2/M checkpoints after exposure to low doses of ionising radiation: implications for hyper-radiosensitivity. DNA Repair (Amst.) 9, 48–57 (2010).

    CAS  Google Scholar 

  171. Mir, S. E. et al. In silico analysis of kinase expression identifies WEE1 as a gatekeeper against mitotic catastrophe in glioblastoma. Cancer Cell 18, 244–257 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  172. Groth, P. et al. Homologous recombination repairs secondary replication induced DNA double-strand breaks after ionizing radiation. Nucleic Acids Res. 40, 6585–6594 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  173. Sauer, R. et al. Preoperative versus postoperative chemoradiotherapy for rectal cancer. N. Engl. J. Med. 351, 1731–1740 (2004).

    CAS  PubMed  Google Scholar 

  174. Sauer, R. et al. Preoperative versus postoperative chemoradiotherapy for locally advanced rectal cancer: results of the German CAO/ARO/AIO-94 randomized phase III trial after a median follow-up of 11 years. J. Clin. Oncol. 30, 1926–1933 (2012).

    CAS  PubMed  Google Scholar 

  175. O'Neill, L. A., Golenbock, D. & Bowie, A. G. The history of Toll-like receptors — redefining innate immunity. Nature Rev. Immunol. 13, 453–460 (2013).

    CAS  Google Scholar 

  176. Pardoll, D. M. The blockade of immune checkpoints in cancer immunotherapy. Nature Rev. Cancer 12, 252–264 (2012).

    CAS  Google Scholar 

  177. Bracci, L., Schiavoni, G., Sistigu, A. & Belardelli, F. Immune-based mechanisms of cytotoxic chemotherapy: implications for the design of novel and rationale-based combined treatments against cancer. Cell Death Differ. 21, 15–25 (2014).

    CAS  PubMed  Google Scholar 

  178. Krysko, O., Løve Aaes, T., Bachert, C., Vandenabeele, P. & Krysko, D. V. Many faces of DAMPs in cancer therapy. Cell Death Dis. 4, e631 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  179. Collins, I. & Garrett, M. D. Targeting the cell division cycle in cancer: CDK and cell cycle checkpoint kinase inhibitors. Curr. Opin. Pharmacol. 5, 366–373 (2005).

    CAS  PubMed  Google Scholar 

  180. Giancotti, F. G. Mechanisms governing metastatic dormancy and reactivation. Cell 155, 750–764 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  181. Banys, M. et al. Dormancy in breast cancer. Breast Cancer (Dove Med. Press) 4, 183–191 (2012).

    CAS  Google Scholar 

  182. Coiras, M., López-Huertas, M. R., Pérez-Olmeda, M. & Alcamí, J. Understanding HIV-1 latency provides clues for the eradication of long-term reservoirs. Nature Rev. Microbiol. 7, 798–812 (2009).

    CAS  Google Scholar 

  183. Komarov, P. G. et al. A chemical inhibitor of p53 that protects mice from the side effects of cancer therapy. Science 285, 1733–1737 (1999).

    CAS  PubMed  Google Scholar 

  184. Gudkov, A. V. & Komarova, E. A. Prospective therapeutic applications of p53 inhibitors. Biochem. Biophys. Res. Commun. 331, 726–736 (2005).

    CAS  PubMed  Google Scholar 

  185. Zeng, X. & King, R. W. An APC/C inhibitor stabilizes cyclin B1 by prematurely terminating ubiquitination. Nature Chem. Biol. 8, 383–392 (2012).

    CAS  Google Scholar 

  186. Felsani, A., Mileo, A. M. & Paggi, M. G. Retinoblastoma family proteins as key targets of the small DNA virus oncoproteins. Oncogene 25, 5277–5285 (2006).

    CAS  PubMed  Google Scholar 

  187. Gong, Y. et al. Pan-cancer genetic analysis identifies PARK2 as a master regulator of G1/S cyclins. Nature Genet. 46, 588–594 (2014).

    CAS  PubMed  Google Scholar 

  188. Nogueira, V. & Hay, N. Molecular pathways: reactive oxygen species homeostasis in cancer cells and implications for cancer therapy. Clin. Cancer Res. 19, 4309–4314 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  190. Trachootham, D., Alexandre, J. & Huang, P. Targeting cancer cells by ROS-mediated mechanisms: a radical therapeutic approach? Nature Rev. Drug Discov. 8, 579–591 (2009).

    CAS  Google Scholar 

  191. Sablina, A. A. et al. The antioxidant function of the p53 tumor suppressor. Nature Med. 11, 1306–1313 (2005).

    CAS  PubMed  Google Scholar 

  192. Sessa, C. et al. Phase I and clinical pharmacological evaluation of aphidicolin glycinate. J. Natl Cancer Inst. 83, 1160–1164 (1991).

    CAS  PubMed  Google Scholar 

  193. Sirbu, B. M. et al. Identification of proteins at active, stalled, and collapsed replication forks using isolation of proteins on nascent DNA (iPOND) coupled with mass spectrometry. J. Biol. Chem. 288, 31458–31467 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  194. López-Contreras, A. J. et al. A proteomic characterization of factors enriched at nascent DNA molecules. Cell Rep. 3, 1105–1116 (2013).

    PubMed  PubMed Central  Google Scholar 

  195. Alabert, C. et al. Nascent chromatin capture proteomics determines chromatin dynamics during DNA replication and identifies unknown fork components. Nature Cell Biol. 16, 281–293 (2014).

    CAS  PubMed  Google Scholar 

  196. Reijns, M. A. et al. Enzymatic removal of ribonucleotides from DNA is essential for mammalian genome integrity and development. Cell 149, 1008–1022 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  Google Scholar 

  198. Knobel, P. A. & Marti, T. M. Translesion DNA synthesis in the context of cancer research. Cancer Cell Int. 11, 39 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  199. Blow, J. J. & Gillespie, P. J. Replication licensing and cancer — a fatal entanglement? Nature Rev. Cancer 8, 799–806 (2008).

    CAS  Google Scholar 

  200. Helmrich, A., Ballarino, M., Nudler, E. & Tora, L. Transcription–replication encounters, consequences and genomic instability. Nature Struct. Mol. Biol. 20, 412–418 (2013).

    CAS  Google Scholar 

  201. Helmrich, A., Ballarino, M. & Tora, L. Collisions between replication and transcription complexes cause common fragile site instability at the longest human genes. Mol. Cell 44, 966–977 (2011).

    CAS  PubMed  Google Scholar 

  202. Barlow, J. H. et al. Identification of early replicating fragile sites that contribute to genome instability. Cell 152, 620–632 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  203. Gilson, E. & Geli, V. How telomeres are replicated. Nature Rev. Mol. Cell Biol. 8, 825–838 (2007).

    CAS  Google Scholar 

  204. Jorgensen, S. et al. The histone methyltransferase SET8 is required for S-phase progression. J. Cell Biol. 179, 1337–1345 (2007).

    PubMed  PubMed Central  Google Scholar 

  205. Tardat, M. et al. The histone H4 Lys 20 methyltransferase PR-Set7 regulates replication origins in mammalian cells. Nature Cell Biol. 12, 1086–1093 (2010).

    CAS  PubMed  Google Scholar 

  206. Jorgensen, S. et al. SET8 is degraded via PCNA-coupled CRL4(CDT2) ubiquitylation in S phase and after UV irradiation. J. Cell Biol. 192, 43–54 (2011).

    PubMed  PubMed Central  Google Scholar 

  207. Centore, R. C. et al. CRL4Cdt2-mediated destruction of the histone methyltransferase Set8 prevents premature chromatin compaction in S phase. Mol. Cell 40, 22–33 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  208. Abbas, T. et al. CRL4Cdt2 regulates cell proliferation and histone gene expression by targeting PR-Set7/Set8 for degradation. Mol. Cell 40, 9–21 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  209. Marechal, A. et al. PRP19 transforms into a sensor of RPA–ssDNA after DNA damage and drives ATR activation via a ubiquitin-mediated circuitry. Mol. Cell 53, 235–246 (2014).

    CAS  PubMed  Google Scholar 

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

    CAS  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  212. Budke, B. et al. RI-1: a chemical inhibitor of RAD51 that disrupts homologous recombination in human cells. Nucleic Acids Res. 40, 7347–7357 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  213. Helleday, T. Putting poly (ADP-ribose) polymerase and other DNA repair inhibitors into clinical practice. Curr. Opin. Oncol. 25, 609–614 (2013).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  215. 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).

    CAS  Google Scholar 

  216. Beck, H. et al. Regulators of cyclin-dependent kinases are crucial for maintaining genome integrity in S phase. J. Cell Biol. 188, 629–638 (2010). References 72, 73 and 216 show that inhibition of the signalling kinase WEE1 enhances replicative stress.

    CAS  PubMed  PubMed Central  Google Scholar 

  217. Li, R., Waga, S., Hannon, G. J., Beach, D. & Stillman, B. Differential effects by the p21 CDK inhibitor on PCNA-dependent DNA replication and repair. Nature 371, 534–537 (1994).

    CAS  PubMed  Google Scholar 

  218. Waga, S., Hannon, G. J., Beach, D. & Stillman, B. The p21 inhibitor of cyclin-dependent kinases controls DNA replication by interaction with PCNA. Nature 369, 574–578 (1994).

    CAS  PubMed  Google Scholar 

  219. Waga, S., Li, R. & Stillman, B. p53-induced p21 controls DNA replication. Leukemia 11 (Suppl. 3), 321–323 (1997).

    PubMed  Google Scholar 

  220. Lu, X., Nannenga, B. & Donehower, L. A. PPM1D dephosphorylates Chk1 and p53 and abrogates cell cycle checkpoints. Genes Dev. 19, 1162–1174 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  221. Gilmartin, A. G. et al. Allosteric Wip1 phosphatase inhibition through flap-subdomain interaction. Nature Chem. Biol. 10, 181–187 (2014).

    CAS  Google Scholar 

  222. Conforti, F., Sayan, A. E., Sreekumar, R. & Sayan, B. S. Regulation of p73 activity by post-translational modifications. Cell Death Dis. 3, e285 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  223. Cannell, I. G. et al. p38 MAPK/MK2-mediated induction of miR-34c following DNA damage prevents Myc-dependent DNA replication. Proc. Natl Acad. Sci. USA 107, 5375–5380 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  224. Damrot, J. et al. DNA replication arrest in response to genotoxic stress provokes early activation of stress-activated protein kinases (SAPK/JNK). J. Mol. Biol. 385, 1409–1421 (2009).

    CAS  PubMed  Google Scholar 

  225. Karin, M., Yamamoto, Y. & Wang, Q. M. The IKK NF-κB system: a treasure trove for drug development. Nature Rev. Drug Discov. 3, 17–26 (2004).

    CAS  Google Scholar 

  226. Hennessy, B. T., Smith, D. L., Ram, P. T., Lu, Y. & Mills, G. B. Exploiting the PI3K/AKT pathway for cancer drug discovery. Nature Rev. Drug Discov. 4, 988–1004 (2005).

    CAS  Google Scholar 

  227. Fulda, S. & Vucic, D. Targeting IAP proteins for therapeutic intervention in cancer. Nature Rev. Drug Discov. 11, 109–124 (2012).

    CAS  Google Scholar 

  228. Cragg, M. S., Harris, C., Strasser, A. & Scott, C. L. Unleashing the power of inhibitors of oncogenic kinases through BH3 mimetics. Nature Rev. Cancer 9, 321–326 (2009).

    CAS  Google Scholar 

  229. Cipolat, S. et al. Mitochondrial rhomboid PARL regulates cytochrome c release during apoptosis via OPA1-dependent cristae remodeling. Cell 126, 163–175 (2006).

    CAS  PubMed  Google Scholar 

  230. Frezza, C. et al. OPA1 controls apoptotic cristae remodeling independently from mitochondrial fusion. Cell 126, 177–189 (2006).

    CAS  PubMed  Google Scholar 

  231. Sato, S. et al. Marine natural product aurilide activates the OPA1-mediated apoptosis by binding to prohibitin. Chem. Biol. 18, 131–139 (2011).

    CAS  PubMed  Google Scholar 

  232. Zong, W. X., Ditsworth, D., Bauer, D. E., Wang, Z. Q. & Thompson, C. B. Alkylating DNA damage stimulates a regulated form of necrotic cell death. Genes Dev. 18, 1272–1282 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  233. Schmitt, C. A. Senescence, apoptosis and therapy — cutting the lifelines of cancer. Nature Rev. Cancer 3, 286–295 (2003).

    CAS  Google Scholar 

  234. Ishikawa, F. Portrait of replication stress viewed from telomeres. Cancer Sci. 104, 790–794 (2013).

    CAS  PubMed  Google Scholar 

  235. Martinez, P. et al. Increased telomere fragility and fusions resulting from TRF1 deficiency lead to degenerative pathologies and increased cancer in mice. Genes Dev. 23, 2060–2075 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  236. McNees, C. J. et al. ATR suppresses telomere fragility and recombination but is dispensable for elongation of short telomeres by telomerase. J. Cell Biol. 188, 639–652 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  237. Mohr, S. E., Smith, J. A., Shamu, C. E., Neumuller, R. A. & Perrimon, N. RNAi screening comes of age: improved techniques and complementary approaches. Nature Rev. Mol. Cell Biol. 15, 591–600 (2014).

    CAS  Google Scholar 

  238. Guzi, T. J. et al. Targeting the replication checkpoint using SCH 900776, a potent and functionally selective CHK1 inhibitor identified via high content screening. Mol. Cancer Ther. 10, 591–602 (2011).

    CAS  PubMed  Google Scholar 

  239. Paulsen, R. D. et al. A genome-wide siRNA screen reveals diverse cellular processes and pathways that mediate genome stability. Mol. Cell 35, 228–239 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  240. Beck, H., Menzel, T., Syljuasen, R. G. & Sorensen, C. S. High-throughput siRNA screens using γH2AX as marker uncover key regulators of genome integrity in mammalian cells. Cell Cycle 9, 2257–2258 (2010).

    CAS  PubMed  Google Scholar 

  241. Fredebohm, J., Wolf, J., Hoheisel, J. D. & Boettcher, M. Depletion of RAD17 sensitizes pancreatic cancer cells to gemcitabine. J. Cell Sci. 126, 3380–3389 (2013).

    CAS  PubMed  Google Scholar 

  242. Azorsa, D. O. et al. Synthetic lethal RNAi screening identifies sensitizing targets for gemcitabine therapy in pancreatic cancer. J. Transl Med. 7, 43 (2009).

    PubMed  PubMed Central  Google Scholar 

  243. Giroux, V., Iovanna, J. & Dagorn, J. C. Probing the human kinome for kinases involved in pancreatic cancer cell survival and gemcitabine resistance. FASEB J. 20, 1982–1991 (2006).

    CAS  PubMed  Google Scholar 

  244. Raman, M., Havens, C. G., Walter, J. C. & Harper, J. W. A genome-wide screen identifies p97 as an essential regulator of DNA damage-dependent CDT1 destruction. Mol. Cell 44, 72–84 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  245. Piwko, W. et al. RNAi-based screening identifies the Mms22L–Nfkbil2 complex as a novel regulator of DNA replication in human cells. EMBO J. 29, 4210–4222 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  246. López-Contreras, A. J., Gutierrez-Martinez, P., Specks, J., Rodrigo-Perez, S. & Fernandez-Capetillo, O. An extra allele of Chk1 limits oncogene-induced replicative stress and promotes transformation. J. Exp. Med. 209, 455–461 (2012).

    PubMed  PubMed Central  Google Scholar 

  247. Schoppy, D. W. et al. Oncogenic stress sensitizes murine cancers to hypomorphic suppression of ATR. J. Clin. Invest. 122, 241–252 (2012).

    CAS  PubMed  Google Scholar 

  248. Murga, M. et al. A mouse model of ATR–Seckel shows embryonic replicative stress and accelerated aging. Nature Genet. 41, 891–898 (2009).

    CAS  PubMed  Google Scholar 

  249. Murga, M. et al. Exploiting oncogene-induced replicative stress for the selective killing of Myc-driven tumors. Nature Struct. Mol. Biol. 18, 1331–1335 (2011).

    CAS  Google Scholar 

  250. Forshew, T. et al. Noninvasive identification and monitoring of cancer mutations by targeted deep sequencing of plasma DNA. Sci. Transl Med. 4, 136ra68 (2012).

    PubMed  Google Scholar 

  251. Vogelstein, B. et al. Cancer genome landscapes. Science 339, 1546–1558 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  252. Barretina, J. et al. The Cancer Cell Line Encyclopedia enables predictive modelling of anticancer drug sensitivity. Nature 483, 603–607 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  254. Gonzalez de Castro, D., Clarke, P. A., Al-Lazikani, B. & Workman, P. Personalized cancer medicine: molecular diagnostics, predictive biomarkers, and drug resistance. Clin. Pharmacol. Ther. 93, 252–259 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  255. Burrell, R. A. et al. Replication stress links structural and numerical cancer chromosomal instability. Nature 494, 492–496 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  256. Tentler, J. J. et al. Patient-derived tumour xenografts as models for oncology drug development. Nature Rev. Clin. Oncol. 9, 338–350 (2012).

    CAS  Google Scholar 

  257. Tuduri, S., Tourriere, H. & Pasero, P. Defining replication origin efficiency using DNA fiber assays. Chromosome Res. 18, 91–102 (2010).

    CAS  PubMed  Google Scholar 

  258. Huberman, J. A. & Riggs, A. D. Autoradiography of chromosomal DNA fibers from Chinese hamster cells. Proc. Natl Acad. Sci. USA 55, 599–606 (1966).

    CAS  PubMed  PubMed Central  Google Scholar 

  259. Jackson, D. A. & Pombo, A. Replicon clusters are stable units of chromosome structure: evidence that nuclear organization contributes to the efficient activation and propagation of S phase in human cells. J. Cell Biol. 140, 1285–1295 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  260. Michalet, X. et al. Dynamic molecular combing: stretching the whole human genome for high-resolution studies. Science 277, 1518–1523 (1997).

    CAS  PubMed  Google Scholar 

  261. Petermann, E. et al. Chk1 requirement for high global rates of replication fork progression during normal vertebrate S phase. Mol. Cell. Biol. 26, 3319–3326 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  262. Leung, K. H., El Hassan, M. A. & Bremner, R. A rapid and efficient method to purify proteins at replication forks under native conditions. Biotechniques 55, 204–206 (2013).

    CAS  PubMed  Google Scholar 

  263. Sirbu, B. M., Couch, F. B. & Cortez, D. Monitoring the spatiotemporal dynamics of proteins at replication forks and in assembled chromatin using isolation of proteins on nascent DNA. Nature Protoc. 7, 594–605 (2012).

    CAS  Google Scholar 

  264. Sirbu, B. M. et al. Analysis of protein dynamics at active, stalled, and collapsed replication forks. Genes Dev. 25, 1320–1327 (2011). References 193, 194, 263 and 264 identify factors that associate with replication forks and are involved in DNA replication (stressed and non-stressed).

    CAS  PubMed  PubMed Central  Google Scholar 

  265. Robison, J. G., Elliott, J., Dixon, K. & Oakley, G. G. Replication protein A and the Mre11•Rad50•Nbs1 complex co-localize and interact at sites of stalled replication forks. J. Biol. Chem. 279, 34802–34810 (2004).

    CAS  PubMed  Google Scholar 

  266. Petermann, E. & Helleday, T. Pathways of mammalian replication fork restart. Nature Rev. Mol. Cell Biol. 11, 683–687 (2010).

    CAS  Google Scholar 

  267. Petermann, E., Orta, M. L., Issaeva, N., Schultz, N. & Helleday, T. Hydroxyurea-stalled replication forks become progressively inactivated and require two different RAD51-mediated pathways for restart and repair. Mol. Cell 37, 492–502 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  268. Ying, S., Hamdy, F. C. & Helleday, T. Mre11-dependent degradation of stalled DNA replication forks is prevented by BRCA2 and PARP1. Cancer Res. 72, 2814–2821 (2012).

    CAS  PubMed  Google Scholar 

  269. Bryant, H. E. et al. PARP is activated at stalled forks to mediate Mre11-dependent replication restart and recombination. EMBO J. 28, 2601–2615 (2009). References 267 and 269 show that replication restarts after fork stalling.

    CAS  PubMed  PubMed Central  Google Scholar 

  270. Mouron, S. et al. Repriming of DNA synthesis at stalled replication forks by human PrimPol. Nature Struct. Mol. Biol. 20, 1383–1389 (2013).

    CAS  Google Scholar 

  271. Bianchi, J. et al. PrimPol bypasses UV photoproducts during eukaryotic chromosomal DNA replication. Mol. Cell 52, 566–573 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  272. Constantinou, A. Rescue of replication failure by Fanconi anaemia proteins. Chromosoma 121, 21–36 (2012).

    CAS  PubMed  Google Scholar 

  273. Sato, K. et al. Histone chaperone activity of Fanconi anemia proteins, FANCD2 and FANCI, is required for DNA crosslink repair. EMBO J. 31, 3524–3536 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  274. Unno, J. et al. FANCD2 binds CtIP and regulates DNA-end resection during DNA interstrand crosslink repair. Cell Rep. 7, 1039–1047 (2014).

    CAS  PubMed  Google Scholar 

  275. Woodward, A. M. et al. Excess Mcm2–7 license dormant origins of replication that can be used under conditions of replicative stress. J. Cell Biol. 173, 673–683 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  276. Ge, X. Q., Jackson, D. A. & Blow, J. J. Dormant origins licensed by excess Mcm2–7 are required for human cells to survive replicative stress. Genes Dev. 21, 3331–3341 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  277. Blow, J. J., Ge, X. Q. & Jackson, D. A. How dormant origins promote complete genome replication. Trends Biochem. Sci. 36, 405–414 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  278. Yekezare, M., Gómez-González, B. & Diffley, J. F. Controlling DNA replication origins in response to DNA damage — inhibit globally, activate locally. J. Cell Sci. 126, 1297–1306 (2013).

    CAS  PubMed  Google Scholar 

  279. Zimmerman, K. M., Jones, R. M., Petermann, E. & Jeggo, P. A. Diminished origin-licensing capacity specifically sensitizes tumor cells to replication stress. Mol. Cancer Res. 11, 370–380 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  280. Toledo, L. I. et al. ATR prohibits replication catastrophe by preventing global exhaustion of RPA. Cell 155, 1088–1103 (2013).

    CAS  PubMed  Google Scholar 

  281. Zou, L. Single- and double-stranded DNA: building a trigger of ATR-mediated DNA damage response. Genes Dev. 21, 879–885 (2007).

    CAS  PubMed  Google Scholar 

  282. Cimprich, K. A. & Cortez, D. ATR: an essential regulator of genome integrity. Nature Rev. Mol. Cell Biol. 9, 616–627 (2008).

    CAS  Google Scholar 

  283. Shiotani, B. & Zou, L. ATR signaling at a glance. J. Cell Sci. 122, 301–304 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  284. Watson, J. D. & Crick, F. H. Molecular structure of nucleic acids; a structure for deoxyribose nucleic acid. Nature 171, 737–738 (1953).

    CAS  PubMed  Google Scholar 

  285. Watson, J. D. & Crick, F. H. Genetical implications of the structure of deoxyribonucleic acid. Nature 171, 964–967 (1953).

    CAS  PubMed  Google Scholar 

  286. Heidelberger, C. et al. Fluorinated pyrimidines, a new class of tumour-inhibitory compounds. Nature 179, 663–666 (1957).

    CAS  PubMed  Google Scholar 

  287. Rosenberg, B., Vancamp, L. & Krigas, T. Inhibition of cell division in Escherichia coli by electrolysis products from a platinum electrode. Nature 205, 698–699 (1965).

    CAS  PubMed  Google Scholar 

  288. Einhorn, L. H. & Donohue, J. Cis-diamminedichloroplatinum, vinblastine, and bleomycin combination chemotherapy in disseminated testicular cancer. Ann. Intern. Med. 87, 293–298 (1977).

    CAS  PubMed  Google Scholar 

  289. Heinemann, V., Hertel, L. W., Grindey, G. B. & Plunkett, W. Comparison of the cellular pharmacokinetics and toxicity of 2′,2′-difluorodeoxycytidine and 1-β-d-arabinofuranosylcytosine. Cancer Res. 48, 4024–4031 (1988).

    CAS  PubMed  Google Scholar 

  290. Paulovich, A. G. & Hartwell, L. H. A checkpoint regulates the rate of progression through S phase in S. cerevisiae in response to DNA damage. Cell 82, 841–847 (1995).

    CAS  PubMed  Google Scholar 

  291. Cimprich, K. A., Shin, T. B., Keith, C. T. & Schreiber, S. L. cDNA cloning and gene mapping of a candidate human cell cycle checkpoint protein. Proc. Natl Acad. Sci. USA 93, 2850–2855 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  292. 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).

    CAS  PubMed  Google Scholar 

  293. Flaggs, G. et al. Atm-dependent interactions of a mammalian Chk1 homolog with meiotic chromosomes. Curr. Biol. 7, 977–986 (1997).

    CAS  PubMed  Google Scholar 

  294. Bryant, H. E. et al. Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature 434, 913–917 (2005). References 269 and 294 highlight the use of PARP inhibitors to delay replication in BRCA2-deficient tumours and enhance replicative stress.

    CAS  PubMed  Google Scholar 

  295. DeVita, V. T. Jr & Chu, E. A history of cancer chemotherapy. Cancer Res. 68, 8643–8653 (2008).

    CAS  PubMed  Google Scholar 

  296. Plunkett, W., Huang, P., Searcy, C. E. & Gandhi, V. Gemcitabine: preclinical pharmacology and mechanisms of action. Semin. Oncol. 23, 3–15 (1996).

    CAS  PubMed  Google Scholar 

  297. Sandhu, S. K. et al. The poly(ADP-ribose) polymerase inhibitor niraparib (MK4827) in BRCA mutation carriers and patients with sporadic cancer: a phase 1 dose-escalation trial. Lancet Oncol. 14, 882–892 (2013).

    CAS  PubMed  Google Scholar 

  298. 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).

    CAS  PubMed  Google Scholar 

  299. 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).

    CAS  PubMed  Google Scholar 

  300. Karp, J. E. et al. Phase I and pharmacologic trial of cytosine arabinoside with the selective checkpoint 1 inhibitor Sch 900776 in refractory acute leukemias. Clin. Cancer Res. 18, 6723–6731 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  301. Fokas, E. et al. Targeting ATR in vivo using the novel inhibitor VE-822 results in selective sensitization of pancreatic tumors to radiation. Cell Death Dis. 3, e441 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  302. Hall, A. B. et al. Potentiation of tumor responses to DNA damaging therapy by the selective ATR inhibitor VX-970. Oncotarget 5, 5674–5685 (2014).

    PubMed  PubMed Central  Google Scholar 

  303. Foote, K. M. et al. Discovery of 4-{4-[(3R)-3- methylmorpholin-4-yl]-6-[1-(methylsulfonyl)cyclopropyl]pyrimidin-2-y l}-1H-indole (AZ20): a potent and selective inhibitor of ATR protein kinase with monotherapy in vivo antitumor activity. J. Med. Chem. 56, 2125–2138 (2013).

    CAS  PubMed  Google Scholar 

  304. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors are indebted to K. Grønbæk, Copenhagen University Hospital, Denmark, for critically reading the manuscript and for many helpful suggestions. M.D. was supported by the Wilhelm Sander Stiftung, the German José Carreras Leukemia Foundation and the Else Kröner–Fresenius–Stiftung to carry out studies on replicative stress. C.S.S. is supported by the Danish Cancer Society, the Danish Medical Research Council, the Novo Nordisk Foundation and Lundbeckfonden.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Matthias Dobbelstein or Claus Storgaard Sørensen.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

FURTHER INFORMATION

FDA

DATABASES

ClinicalTrials.gov

PowerPoint slides

Glossary

Checkpoints

Signalling events during the cell cycle that prevent further progression.

Cyclin-dependent kinase

(CDK). A class of kinase that associates with partner proteins known as cyclins. Specific CDKs are active at various phases of the cell cycle to promote cell cycle progression.

Replicative stress

The perturbation of DNA replication that interferes with timely and error-free completion of S phase.

Nucleoside analogues

Compounds with similarity to nucleosides (components of DNA or RNA) that are often used as drugs to interfere with the polymerization of (deoxy)ribonucleotides.

Checkpoint kinase

A protein kinase that is activated by stress signals and halts cell cycle progression.

Platinum compounds

A class of chemotherapeutics that include cisplatin, oxaliplatin and carboplatin. Platinum compounds form intrastrand and interstrand crosslinks on DNA.

Translesion synthesis

DNA synthesis by specific enzymes — DNA polymerase-ζ and DNA polymerase-η — to bypass small lesions in the template strand.

Topoisomerase inhibitors

A class of drugs that bind to topoisomerase I (rotating DNA around a nick to relax DNA supercoiling) or topoisomerase II (passing one portion of DNA through another). These inhibitors typically associate with the topoisomerases while bound to DNA, which creates a barrier to DNA replication.

Poly(ADP-ribose) polymerase

(PARP). A class of enzyme that is involved in facilitating DNA repair.

Isolation of proteins on nascent DNA

(iPOND). A technology that involves the labelling of newly synthesized DNA and subsequent purification of proteins that associate with this DNA.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dobbelstein, M., Sørensen, C. Exploiting replicative stress to treat cancer. Nat Rev Drug Discov 14, 405–423 (2015). https://doi.org/10.1038/nrd4553

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrd4553

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer