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:

The impact of replication stress on replication dynamics and DNA damage in vertebrate cells

Key Points

  • The progression of replication forks is impaired when they encounter intrinsically hard-to-replicate sequences (such as repeats and secondary structures) or the transcription machinery, or when DNA precursors become limiting.

  • Upon fork slowing, cells maintain their replication rate by adjusting the density of initiation events to fork speed, and vice versa. This regulatory process is known as compensation.

  • Fork slowing becomes pathological when it reaches the threshold that exhausts the pool of latent origins, resulting in compensation failure.

  • Activation of the ATM pathway by DNA lesions elicits moderate fork slowing and the compensatory activation of latent origins.

  • The activation of the DNA damage checkpoint both facilitates repair and helps replication to proceed across intrinsically hard to replicate or damaged sequences upon firing of extra-origins.

  • Oncogene overexpression promotes reactive oxygen species (ROS) overproduction and upregulation of cyclin-dependent kinase (CDK), resulting in DNA damage, unscheduled activation of structure-dependent nucleases, precursor shortage and subsequent genome instability.

Abstract

The interplay between replication stress and the S phase checkpoint is a key determinant of genome maintenance, and has a major impact on human diseases, notably, tumour initiation and progression. Recent studies have yielded insights into sequence-dependent and sequence-independent sources of endogenous replication stress. These stresses result in nuclease-induced DNA damage, checkpoint activation and genome-wide replication fork slowing. Several hypotheses have been proposed to account for the mechanisms involved in this complex response. Recent results have shown that the slowing of the replication forks most commonly results from DNA precursor starvation. By concomitantly increasing the density of replication initiation, the cell elicits an efficient compensatory strategy to avoid mitotic anomalies and the inheritance of damage over cell generations.

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: Consequences of perturbations of fork progression.
Figure 2: Hard-to-replicate sequences.
Figure 3: Impact of CHK1 or WEE1 deficiency on replication dynamics.

Similar content being viewed by others

References

  1. Tourriere, H. & Pasero, P. Maintenance of fork integrity at damaged DNA and natural pause sites. DNA Repair (Amst.) 6, 900–913 (2007).

    Article  CAS  Google Scholar 

  2. Macheret, M. & Halazonetis, T. D. DNA replication stress as a hallmark of cancer. Annu. Rev. Pathol. 10, 425–448 (2015).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Halazonetis, T. D., Gorgoulis, V. G. & Bartek, J. An oncogene-induced DNA damage model for cancer development. Science 319, 1352–1355 (2008). The authors propose for the first time that oncogene activation in precancerous lesions induces replication stress, resulting in the formation of DNA breaks and subsequent genomic instability. This ultimately selects for cells that are deficient in checkpoint surveillance.

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  6. Dereli-Oz, A., Versini, G. & Halazonetis, T. D. Studies of genomic copy number changes in human cancers reveal signatures of DNA replication stress. Mol. Oncol. 5, 308–314 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Le Tallec, B. et al. Common fragile site profiling in epithelial and erythroid cells reveals that most recurrent cancer deletions lie in fragile sites hosting large genes. Cell Rep. 4, 420–428 (2013).

    Article  CAS  PubMed  Google Scholar 

  9. Aladjem, M. I. & Redon, C. E. Order from clutter: selective interactions at mammalian replication origins. Nat. Rev. Genet. 18, 101–116 (2017).

    Article  CAS  PubMed  Google Scholar 

  10. 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).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  12. Dehe, P. M. & Gaillard, P. H. Control of structure-specific endonucleases to maintain genome stability. Nat. Rev. Mol. Cell Biol. 18, 315–330 (2017).

    Article  CAS  PubMed  Google Scholar 

  13. Gilbert, D. M. et al. Space and time in the nucleus: developmental control of replication timing and chromosome architecture. Cold Spring Harb. Symp. Quant. Biol. 75, 143–153 (2010).

    Article  CAS  PubMed  Google Scholar 

  14. Gilbert, D. M. Cell fate transitions and the replication timing decision point. J. Cell Biol. 191, 899–903 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Pope, B. D. et al. Replication-timing boundaries facilitate cell-type and species-specific regulation of a rearranged human chromosome in mouse. Hum. Mol. Genet. 21, 4162–4170 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Dileep, V., Rivera-Mulia, J. C., Sima, J. & Gilbert, D. M. Large-scale chromatin structure-function relationships during the cell cycle and development: insights from replication timing. Cold Spring Harb. Symp. Quant. Biol. 80, 53–63 (2015).

    Article  PubMed  Google Scholar 

  17. Dileep, V. et al. Topologically associating domains and their long-range contacts are established during early G1 coincident with the establishment of the replication-timing program. Genome Res. 25, 1104–1113 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Sexton, T. & Cavalli, G. The role of chromosome domains in shaping the functional genome. Cell 160, 1049–1059 (2015).

    Article  CAS  PubMed  Google Scholar 

  19. Naumova, N. et al. Organization of the mitotic chromosome. Science 342, 948–953 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Chagin, V. O. et al. 4D visualization of replication foci in mammalian cells corresponding to individual replicons. Nat. Commun. 7, 11231 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Schepers, A. & Papior, P. Why are we where we are? Understanding replication origins and initiation sites in eukaryotes using ChIP-approaches. Chromosome Res. 18, 63–77 (2010).

    Article  CAS  PubMed  Google Scholar 

  22. Deegan, T. D. & Diffley, J. F. MCM: one ring to rule them all. Curr. Opin. Struct. Biol. 37, 145–151 (2016).

    Article  CAS  PubMed  Google Scholar 

  23. Gambus, A., Khoudoli, G. A., Jones, R. C. & Blow, J. J. MCM2-7 form double hexamers at licensed origins in Xenopus egg extract. J. Biol. Chem. 286, 11855–11864 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Evrin, C. et al. A double-hexameric MCM2-7 complex is loaded onto origin DNA during licensing of eukaryotic DNA replication. Proc. Natl Acad. Sci. USA 106, 20240–20245 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Siddiqui, K., On, K. F. & Diffley, J. F. Regulating DNA replication in eukarya. Cold Spring Harb. Perspect. Biol. 5, a012930 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Yeeles, J. T., Deegan, T. D., Janska, A., Early, A. & Diffley, J. F. Regulated eukaryotic DNA replication origin firing with purified proteins. Nature 519, 431–435 (2015). This pioneer work reports the purification and in vitro reconstitution of the different steps of pre-RC assembly and activation up to origin firing. The authors identified for the first time the minimal set of proteins essential for initiating replication in eukaryotes.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Hyrien, O. Peaks cloaked in the mist: the landscape of mammalian replication origins. J. Cell Biol. 208, 147–160 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Prioleau, M. N. & MacAlpine, D. M. DNA replication origins-where do we begin? Genes Dev. 30, 1683–1697 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Fragkos, M., Ganier, O., Coulombe, P. & Mechali, M. DNA replication origin activation in space and time. Nat. Rev. Mol. Cell Biol. 16, 360–374 (2015).

    Article  CAS  PubMed  Google Scholar 

  30. Lujan, S. A., Williams, J. S. & Kunkel, T. A. DNA polymerases divide the labor of genome replication. Trends Cell Biol. 26, 640–654 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Yeeles, J. T., Janska, A., Early, A. & Diffley, J. F. How the eukaryotic replisome achieves rapid and efficient DNA replication. Mol. Cell 65, 105–116 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Kurat, C. F., Yeeles, J. T., Patel, H., Early, A. & Diffley, J. F. Chromatin controls DNA replication origin selection, lagging-strand synthesis, and replication fork rates. Mol. Cell 65, 117–130 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Devbhandari, S., Jiang, J., Kumar, C., Whitehouse, I. & Remus, D. Chromatin constrains the initiation and elongation of DNA replication. Mol. Cell 65, 131–141 (2017).

    Article  CAS  PubMed  Google Scholar 

  34. Guilbaud, G. et al. Evidence for sequential and increasing activation of replication origins along replication timing gradients in the human genome. PLoS Comput. Biol. 7, e1002322 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Techer, H. et al. Replication dynamics: biases and robustness of DNA fiber analysis. J. Mol. Biol. 425, 4845–4855 (2013).

    Article  CAS  PubMed  Google Scholar 

  36. Koundrioukoff, S. et al. Stepwise activation of the ATR signaling pathway upon increasing replication stress impacts fragile site integrity. PLoS Genet. 9, e1003643 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Speroni, J., Federico, M. B., Mansilla, S. F., Soria, G. & Gottifredi, V. Kinase-independent function of checkpoint kinase 1 (Chk1) in the replication of damaged DNA. Proc. Natl Acad. Sci. USA 109, 7344–7349 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Tuduri, S. et al. Topoisomerase I suppresses genomic instability by preventing interference between replication and transcription. Nat. Cell Biol. 11, 1315–1324 (2009). This study shows how topoisomerase I deficiency leads to genomic instability through enhanced conflicts between replication forks and the transcription machinery.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Poli, J. et al. dNTP pools determine fork progression and origin usage under replication stress. EMBO J. 31, 883–894 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Techer, H. et al. Signaling from Mus81-Eme2-dependent DNA damage elicited by Chk1 deficiency modulates replication fork speed and origin usage. Cell Rep. 14, 1114–1127 (2016). This study shows that Chk1 deficiency triggers nuclease-dependent DNA damage. Damage activates the ATM pathway and, in turn, induces fork slowing and subsequent firing of latent origins, helping replication to proceed along damaged templates.

    Article  CAS  PubMed  Google Scholar 

  41. Wilhelm, T. et al. Slow replication fork velocity of homologous recombination-defective cells results from endogenous oxidative stress. PLoS Genet. 12, e1006007 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Chabosseau, P. et al. Pyrimidine pool imbalance induced by BLM helicase deficiency contributes to genetic instability in Bloom syndrome. Nat. Commun. 2, 368 (2011).

    Article  PubMed  CAS  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Gay, S. et al. Nucleotide supply, not local histone acetylation, sets replication origin usage in transcribed regions. EMBO Rep. 11, 698–704 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Groth, A. et al. Regulation of replication fork progression through histone supply and demand. Science 318, 1928–1931 (2007).

    Article  CAS  PubMed  Google Scholar 

  46. Mejlvang, J. et al. New histone supply regulates replication fork speed and PCNA unloading. J. Cell Biol. 204, 29–43 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Dungrawala, H. et al. The replication checkpoint prevents two types of fork collapse without regulating replisome stability. Mol. Cell 59, 998–1010 (2015). This study provides an exhaustive and quantitative analysis of the protein landscape at replication forks in a variety of experimental conditions, which reveals unexpected mechanisms by which the DDR preserves genome integrity.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Gordenin, D. A. & Resnick, M. A. Yeast ARMs (DNA at-risk motifs) can reveal sources of genome instability. Mutat. Res. 400, 45–58 (1998).

    Article  CAS  PubMed  Google Scholar 

  49. Liu, P., Carvalho, C. M., Hastings, P. J. & Lupski, J. R. Mechanisms for recurrent and complex human genomic rearrangements. Curr. Opin. Genet. Dev. 22, 211–220 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Sutherland, G. R., Baker, E. & Richards, R. I. Fragile sites still breaking. Trends Genet. 14, 501–506 (1998).

    Article  CAS  PubMed  Google Scholar 

  51. Mirkin, E. V. & Mirkin, S. M. Replication fork stalling at natural impediments. Microbiol. Mol. Biol. Rev. 71, 13–35 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Wang, G. & Vasquez, K. M. Impact of alternative DNA structures on DNA damage, DNA repair, and genetic instability. DNA Repair (Amst.) 19, 143–151 (2014).

    Article  CAS  Google Scholar 

  53. Usdin, K., House, N. C. & Freudenreich, C. H. Repeat instability during DNA repair: insights from model systems. Crit. Rev. Biochem. Mol. Biol. 50, 142–167 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Wickramasinghe, C. M., Arzouk, H., Frey, A., Maiter, A. & Sale, J. E. Contributions of the specialised DNA polymerases to replication of structured DNA. DNA Repair (Amst.) 29, 83–90 (2015).

    Article  CAS  Google Scholar 

  55. Barthelemy, J., Hanenberg, H. & Leffak, M. FANCJ is essential to maintain microsatellite structure genome-wide during replication stress. Nucleic Acids Res. 44, 6803–6816 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Mendoza, O., Bourdoncle, A., Boule, J. B., Brosh, R. M. Jr & Mergny, J. L. G-Quadruplexes and helicases. Nucleic Acids Res. 44, 1989–2006 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Leon-Ortiz, A. M., Svendsen, J. & Boulton, S. J. Metabolism of DNA secondary structures at the eukaryotic replication fork. DNA Repair (Amst.) 19, 152–162 (2014).

    Article  CAS  Google Scholar 

  58. Schmidt, M. H. & Pearson, C. E. Disease-associated repeat instability and mismatch repair. DNA Repair (Amst.) 38, 117–126 (2016).

    Article  CAS  Google Scholar 

  59. Maizels, N. G4-associated human diseases. EMBO Rep. 16, 910–922 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Rhodes, D. & Lipps, H. J. G-Quadruplexes and their regulatory roles in biology. Nucleic Acids Res. 43, 8627–8637 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Sollier, J. & Cimprich, K. A. Breaking bad: R-loops and genome integrity. Trends Cell Biol. 25, 514–522 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Santos-Pereira, J. M. & Aguilera, A. R loops: new modulators of genome dynamics and function. Nat. Rev. Genet. 16, 583–597 (2015).

    Article  CAS  PubMed  Google Scholar 

  63. Garcia-Muse, T. & Aguilera, A. Transcription-replication conflicts: how they occur and how they are resolved. Nat. Rev. Mol. Cell Biol. 17, 553–563 (2016).

    Article  CAS  PubMed  Google Scholar 

  64. Sollier, J. et al. Transcription-coupled nucleotide excision repair factors promote R-loop-induced genome instability. Mol. Cell 56, 777–785 (2014). This study shows that R-loops are actively processed into DNA double-strand breaks by the NER endonucleases XPF–ERCC1 and XPG.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Stork, C. T. et al. Co-transcriptional R-loops are the main cause of estrogen-induced DNA damage. eLife 5, e17548 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  66. Ginno, P. A., Lott, P. L., Christensen, H. C., Korf, I. & Chedin, F. R-Loop formation is a distinctive characteristic of unmethylated human CpG island promoters. Mol. Cell 45, 814–825 (2012). This study reports the development of DRIP (DNA–RNA hybrid immunoprecipitation), a technique that allows the genome-wide mapping of these hybrids. The authors show that long R-loop structures form preferentially upon the transcription of regions with high GC skew.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Ginno, P. A., Lim, Y. W., Lott, P. L., Korf, I. & Chedin, F. GC skew at the 5′ and 3′ ends of human genes links R-loop formation to epigenetic regulation and transcription termination. Genome Res. 23, 1590–1600 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Sanz, L. A. et al. Prevalent, dynamic, and conserved R-loop structures associate with specific epigenomic signatures in mammals. Mol. Cell 63, 167–178 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Kantidakis, T. et al. Mutation of cancer driver MLL2 results in transcription stress and genome instability. Genes Dev. 30, 408–420 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Pommier, Y., Sun, Y., Huang, S. N. & Nitiss, J. L. Roles of eukaryotic topoisomerases in transcription, replication and genomic stability. Nat. Rev. Mol. Cell Biol. 17, 703–721 (2016).

    Article  CAS  PubMed  Google Scholar 

  71. Bhatia, V. et al. BRCA2 prevents R-loop accumulation and associates with TREX-2 mRNA export factor PCID2. Nature 511, 362–365 (2014). This study shows that BRCA2 and FANC proteins protect genome stability by promoting the elimination of R-loops, which block replication fork progression.

    Article  CAS  PubMed  Google Scholar 

  72. Garcia-Rubio, M. L. et al. The Fanconi anemia pathway protects genome integrity from R-loops. PLoS Genet. 11, e1005674 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  73. Yeo, C. Q. et al. p53 maintains genomic stability by preventing interference between transcription and replication. Cell Rep. 15, 132–146 (2016).

    Article  CAS  PubMed  Google Scholar 

  74. Herrera-Moyano, E., Mergui, X., Garcia-Rubio, M. L., Barroso, S. & Aguilera, A. The yeast and human FACT chromatin-reorganizing complexes solve R-loop-mediated transcription-replication conflicts. Genes Dev. 28, 735–748 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Gan, W. et al. R-Loop-mediated genomic instability is caused by impairment of replication fork progression. Genes Dev. 25, 2041–2056 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Durkin, S. G. & Glover, T. W. Chromosome fragile sites. Annu. Rev. Genet. 41, 169–192 (2007).

    Article  CAS  PubMed  Google Scholar 

  77. Debatisse, M., Le Tallec, B., Letessier, A., Dutrillaux, B. & Brison, O. Common fragile sites: mechanisms of instability revisited. Trends Genet. 28, 22–32 (2012).

    Article  CAS  PubMed  Google Scholar 

  78. 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).

    Article  CAS  PubMed  Google Scholar 

  79. Wilson, T. E. et al. Large transcription units unify copy number variants and common fragile sites arising under replication stress. Genome Res. 25, 189–200 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Letessier, A. et al. Cell-type-specific replication initiation programs set fragility of the FRA3B fragile site. Nature 470, 120–123 (2011).

    Article  CAS  PubMed  Google Scholar 

  81. Le Tallec, B. et al. Molecular profiling of common fragile sites in human fibroblasts. Nat. Struct. Mol. Biol. 18, 1421–1423 (2011).

    Article  CAS  PubMed  Google Scholar 

  82. Gros, J. et al. Post-licensing specification of eukaryotic replication origins by facilitated Mcm2-7 sliding along DNA. Mol. Cell 60, 797–807 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Powell, S. K. et al. Dynamic loading and redistribution of the Mcm2-7 helicase complex through the cell cycle. EMBO J. 34, 531–543 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Chon, H. et al. RNase H2 roles in genome integrity revealed by unlinking its activities. Nucleic Acids Res. 41, 3130–3143 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Madireddy, A. et al. FANCD2 facilitates replication through common fragile sites. Mol. Cell 64, 388–404 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Saponaro, M. et al. RECQL5 controls transcript elongation and suppresses genome instability associated with transcription stress. Cell 157, 1037–1049 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Stillman, B. Deoxynucleoside triphosphate (dNTP) synthesis and destruction regulate the replication of both cell and virus genomes. Proc. Natl Acad. Sci. USA 110, 14120–14121 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Murthy, S. & Reddy, G. P. Replitase: complete machinery for DNA synthesis. J. Cell. Physiol. 209, 711–717 (2006).

    Article  CAS  PubMed  Google Scholar 

  89. Mathews, C. K. Deoxyribonucleotide metabolism, mutagenesis and cancer. Nat. Rev. Cancer 15, 528–539 (2015).

    Article  CAS  PubMed  Google Scholar 

  90. Chabes, A. L., Pfleger, C. M., Kirschner, M. W. & Thelander, L. Mouse ribonucleotide reductase R2 protein: a new target for anaphase-promoting complex-Cdh1-mediated proteolysis. Proc. Natl Acad. Sci. USA 100, 3925–3929 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. D'Angiolella, V. et al. Cyclin F-mediated degradation of ribonucleotide reductase M2 controls genome integrity and DNA repair. Cell 149, 1023–1034 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Hu, C. M. et al. Tumor cells require thymidylate kinase to prevent dUTP incorporation during DNA repair. Cancer Cell 22, 36–50 (2012).

    Article  CAS  PubMed  Google Scholar 

  93. Chen, C. W. et al. The impact of dUTPase on ribonucleotide reductase-induced genome instability in cancer cells. Cell Rep. 16, 1287–1299 (2016).

    Article  CAS  PubMed  Google Scholar 

  94. Hakansson, P., Hofer, A. & Thelander, L. Regulation of mammalian ribonucleotide reduction and dNTP pools after DNA damage and in resting cells. J. Biol. Chem. 281, 7834–7841 (2006).

    Article  CAS  PubMed  Google Scholar 

  95. Pontarin, G. et al. p53R2-dependent ribonucleotide reduction provides deoxyribonucleotides in quiescent human fibroblasts in the absence of induced DNA damage. J. Biol. Chem. 282, 16820–16828 (2007).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  97. Chang, L. et al. ATM-mediated serine 72 phosphorylation stabilizes ribonucleotide reductase small subunit p53R2 protein against MDM2 to DNA damage. Proc. Natl Acad. Sci. USA 105, 18519–18524 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Lin, Z. P., Belcourt, M. F., Cory, J. G. & Sartorelli, A. C. Stable suppression of the R2 subunit of ribonucleotide reductase by R2-targeted short interference RNA sensitizes p53−/− HCT-116 colon cancer cells to DNA-damaging agents and ribonucleotide reductase inhibitors. J. Biol. Chem. 279, 27030–27038 (2004).

    Article  CAS  PubMed  Google Scholar 

  99. Reddy, G. P. & Mathews, C. K. Functional compartmentation of DNA precursors in T4 phage-infected bacteria. J. Biol. Chem. 253, 3461–3467 (1978).

    Article  CAS  PubMed  Google Scholar 

  100. Mathews, C. K. & Sinha, N. K. Are DNA precursors concentrated at replication sites? Proc. Natl Acad. Sci. USA 79, 302–306 (1982).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Franzolin, E. et al. The deoxynucleotide triphosphohydrolase SAMHD1 is a major regulator of DNA precursor pools in mammalian cells. Proc. Natl Acad. Sci. USA 110, 14272–14277 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Clifford, R. et al. SAMHD1 is mutated recurrently in chronic lymphocytic leukemia and is involved in response to DNA damage. Blood 123, 1021–1031 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Bessman, M. J. et al. Enzymatic synthesis of deoxyribonucleic acid. III. The incorporation of pyrimidine and purine analogues into deoxyribonucleic acid. Proc. Natl Acad. Sci. USA 44, 633–640 (1958).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Nick McElhinny, S. A. et al. Genome instability due to ribonucleotide incorporation into DNA. Nat. Chem. Biol. 6, 774–781 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Spadari, S. et al. Control of cell division by aphidicolin without adverse effects upon resting cells. Arzneimittelforschung 35, 1108–1116 (1985).

    CAS  PubMed  Google Scholar 

  107. Dominguez-Kelly, R. et al. Wee1 controls genomic stability during replication by regulating the Mus81-Eme1 endonuclease. J. Cell Biol. 194, 567–579 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Wilhelm, T. et al. Spontaneous slow replication fork progression elicits mitosis alterations in homologous recombination-deficient mammalian cells. Proc. Natl Acad. Sci. USA 111, 763–768 (2014).

    Article  CAS  PubMed  Google Scholar 

  109. Lossaint, G. et al. FANCD2 binds MCM proteins and controls replisome function upon activation of s phase checkpoint signaling. Mol. Cell 51, 678–690 (2013). This study shows that the ATR–FANCD2–FANCI pathway actively regulates fork progression in response to replication stress through a mechanism involving the MCM complex.

    Article  CAS  PubMed  Google Scholar 

  110. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Taylor, J. H. Increase in DNA replication sites in cells held at the beginning of S phase. Chromosoma 62, 291–300 (1977). This pioneer work shows that slowing the progression of replication forks triggers a compensatory increase in the density of initiation events.

    Article  CAS  PubMed  Google Scholar 

  112. Anglana, M., Apiou, F., Bensimon, A. & Debatisse, M. Dynamics of DNA replication in mammalian somatic cells: nucleotide pool modulates origin choice and interorigin spacing. Cell 114, 385–394 (2003).

    Article  CAS  PubMed  Google Scholar 

  113. Marheineke, K. & Hyrien, O. Control of replication origin density and firing time in Xenopus egg extracts: role of a caffeine-sensitive, ATR-dependent checkpoint. J. Biol. Chem. 279, 28071–28081 (2004).

    Article  CAS  PubMed  Google Scholar 

  114. Courbet, S. et al. Replication fork movement sets chromatin loop size and origin choice in mammalian cells. Nature 455, 557–560 (2008).

    Article  CAS  PubMed  Google Scholar 

  115. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Ibarra, A., Schwob, E. & Mendez, J. Excess MCM proteins protect human cells from replicative stress by licensing backup origins of replication. Proc. Natl Acad. Sci. USA 105, 8956–8961 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Kohler, C. et al. Cdc45 is limiting for replication initiation in humans. Cell Cycle 15, 974–985 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  118. Lachaud, C. et al. Ubiquitinated Fancd2 recruits Fan1 to stalled replication forks to prevent genome instability. Science 351, 846–849 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Chen, Y. H. et al. ATR-mediated phosphorylation of FANCI regulates dormant origin firing in response to replication stress. Mol. Cell 58, 323–338 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Sheu, Y. J., Kinney, J. B., Lengronne, A., Pasero, P. & Stillman, B. Domain within the helicase subunit Mcm4 integrates multiple kinase signals to control DNA replication initiation and fork progression. Proc. Natl Acad. Sci. USA 111, E1899–E1908 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Dimitrova, D. S. & Gilbert, D. M. Temporally coordinated assembly and disassembly of replication factories in the absence of DNA synthesis. Nat. Cell Biol. 2, 686–694 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Zachos, G., Rainey, M. D. & Gillespie, D. A. Chk1-deficient tumour cells are viable but exhibit multiple checkpoint and survival defects. EMBO J. 22, 713–723 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Zegerman, P. & Diffley, J. F. DNA replication as a target of the DNA damage checkpoint. DNA Repair (Amst.) 8, 1077–1088 (2009).

    Article  CAS  Google Scholar 

  124. Ge, X. Q. & Blow, J. J. Chk1 inhibits replication factory activation but allows dormant origin firing in existing factories. J. Cell Biol. 191, 1285–1297 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Rybak, P., Waligorska, A., Bujnowicz, L., Hoang, A. & Dobrucki, J. W. Activation of new replication foci under conditions of replication stress. Cell Cycle 14, 2634–2647 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Eykelenboom, J. K. et al. ATR activates the S-M checkpoint during unperturbed growth to ensure sufficient replication prior to mitotic onset. Cell Rep. 5, 1095–1107 (2013).

    Article  CAS  PubMed  Google Scholar 

  127. Buisson, R., Boisvert, J. L., Benes, C. H. & Zou, L. Distinct but concerted roles of ATR, DNA-PK, and Chk1 in countering replication stress during S phase. Mol. Cell 59, 1011–1024 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Maya-Mendoza, A., Petermann, E., Gillespie, D. A., Caldecott, K. W. & Jackson, D. A. Chk1 regulates the density of active replication origins during the vertebrate S phase. EMBO J. 26, 2719–2731 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Katsuno, Y. et al. Cyclin A-Cdk1 regulates the origin firing program in mammalian cells. Proc. Natl Acad. Sci. USA 106, 3184–3189 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Petermann, E., Helleday, T. & Caldecott, K. W. Claspin promotes normal replication fork rates in human cells. Mol. Biol. Cell 19, 2373–2378 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Miotto, B. et al. The RBBP6/ZBTB38/MCM10 axis regulates DNA replication and common fragile site stability. Cell Rep. 7, 575–587 (2014).

    Article  CAS  PubMed  Google Scholar 

  136. Choi, H. J. et al. NEK8 links the ATR-regulated replication stress response and S phase CDK activity to renal ciliopathies. Mol. Cell 51, 423–439 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Fu, H. et al. The DNA repair endonuclease Mus81 facilitates fast DNA replication in the absence of exogenous damage. Nat. Commun. 6, 6746 (2015).

    Article  CAS  PubMed  Google Scholar 

  138. Giunco, S. et al. Cross talk between EBV and telomerase: the role of TERT and NOTCH2 in the switch of latent/lytic cycle of the virus. Cell Death Dis. 6, e1774 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Park, Y. B. et al. Alterations in the INK4a/ARF locus and their effects on the growth of human osteosarcoma cell lines. Cancer Genet. Cytogenet. 133, 105–111 (2002).

    Article  CAS  PubMed  Google Scholar 

  140. Ahuja, A. K. et al. A short G1 phase imposes constitutive replication stress and fork remodelling in mouse embryonic stem cells. Nat. Commun. 7, 10660 (2016). This study shows the massive accumulation of ssDNA gaps, reduced fork speed and frequent fork reversal in mouse embryonic stem cells. These marks of replication stress can be suppressed by delaying the G1–S transition, showing that rapid cell cycle progression challenges genome integrity.

  141. Forment, J. V., Blasius, M., Guerini, I. & Jackson, S. P. Structure-specific DNA endonuclease Mus81/Eme1 generates DNA damage caused by Chk1 inactivation. PLoS ONE 6, e23517 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Murfuni, I. et al. Survival of the replication checkpoint deficient cells requires MUS81-RAD52 function. PLoS Genet. 9, e1003910 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  143. Thompson, R., Montano, R. & Eastman, A. The Mre11 nuclease is critical for the sensitivity of cells to Chk1 inhibition. PLoS ONE 7, e44021 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Pepe, A. & West, S. C. MUS81-EME2 promotes replication fork restart. Cell Rep. 7, 1048–1055 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Buis, J., Stoneham, T., Spehalski, E. & Ferguson, D. O. Mre11 regulates CtIP-dependent double-strand break repair by interaction with CDK2. Nat. Struct. Mol. Biol. 19, 246–252 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. 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  PubMed  PubMed Central  Google Scholar 

  148. Peterson, S. E. et al. Cdk1 uncouples CtIP-dependent resection and Rad51 filament formation during M-phase double-strand break repair. J. Cell Biol. 194, 705–720 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Dehe, P. M. et al. Regulation of Mus81-Eme1 Holliday junction resolvase in response to DNA damage. Nat. Struct. Mol. Biol. 20, 598–603 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Falck, J. et al. CDK targeting of NBS1 promotes DNA-end resection, replication restart and homologous recombination. EMBO Rep. 13, 561–568 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Shechter, D., Costanzo, V. & Gautier, J. Regulation of DNA replication by ATR: signaling in response to DNA intermediates. DNA Repair (Amst.) 3, 901–908 (2004).

    Article  CAS  Google Scholar 

  152. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Niida, H. et al. Essential role of Tip60-dependent recruitment of ribonucleotide reductase at DNA damage sites in DNA repair during G1 phase. Genes Dev. 24, 333–338 (2010). This study shows that, during the G1 phase, RNR re-localizes to repair foci in a TIP60-dependent manner, suggesting that a local increase in dNTP pools favours repair synthesis.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Neelsen, K. J. & Lopes, M. Replication fork reversal in eukaryotes: from dead end to dynamic response. Nat. Rev. Mol. Cell Biol. 16, 207–220 (2015).

    Article  CAS  PubMed  Google Scholar 

  155. Kolinjivadi, A. M. et al. Moonlighting at replication forks — a new life for homologous recombination proteins BRCA1, BRCA2 and RAD51. FEBS Lett. 591, 1083–1100 (2017).

    Article  CAS  PubMed  Google Scholar 

  156. Toledo, L. I. et al. ATR prohibits replication catastrophe by preventing global exhaustion of RPA. Cell 155, 1088–1103 (2013). This study shows that RPA is present in limiting amount in cells, which compromises ssDNA protection upon stringent replication stress, especially in cells deficient in ATR.

    Article  CAS  PubMed  Google Scholar 

  157. Hossain, M. & Stillman, B. Opposing roles for DNA replication initiator proteins ORC1 and CDC6 in control of Cyclin E gene transcription. eLife 5, e12785 (2016). This study shows that ORC1 and CDC6 have opposing effects in controlling the level of cyclin E expression. Cyclin E level then regulates cell passage through the restriction point and S phase onset, which ensures genome stability.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  158. Sanjiv, K. et al. Cancer-specific synthetic lethality between ATR and CHK1 kinase activities. Cell Rep. 14, 298–309 (2016).

    Article  CAS  PubMed  Google Scholar 

  159. Mokrani-Benhelli, H. et al. Primary microcephaly, impaired DNA replication, and genomic instability caused by compound heterozygous ATR mutations. Hum. Mutat. 34, 374–384 (2013).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Lopez-Contreras, A. J. et al. Increased Rrm2 gene dosage reduces fragile site breakage and prolongs survival of ATR mutant mice. Genes Dev. 29, 690–695 (2015). This study shows that extra-alleles of the RNR regulatory subunit RRM2 enhance RNR activity, resulting in reduced chromosomal breaks at fragile sites, and extending the lifespan of ATR-mutant mice.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Hills, S. A. & Diffley, J. F. DNA replication and oncogene-induced replicative stress. Curr. Biol. 24, R435–R444 (2014).

    Article  CAS  PubMed  Google Scholar 

  163. Neelsen, K. J., Zanini, I. M., Herrador, R. & Lopes, M. Oncogenes induce genotoxic stress by mitotic processing of unusual replication intermediates. J. Cell Biol. 200, 699–708 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Maya-Mendoza, A. et al. Myc and Ras oncogenes engage different energy metabolism programs and evoke distinct patterns of oxidative and DNA replication stress. Mol. Oncol. 9, 601–616 (2015).

    Article  CAS  PubMed  Google Scholar 

  165. Srinivasan, S. V., Dominguez-Sola, D., Wang, L. C., Hyrien, O. & Gautier, J. Cdc45 is a critical effector of myc-dependent DNA replication stress. Cell Rep. 3, 1629–1639 (2013).

    Article  CAS  PubMed  Google Scholar 

  166. Jones, R. M. et al. Increased replication initiation and conflicts with transcription underlie Cyclin E-induced replication stress. Oncogene 32, 3744–3753 (2013).

    Article  CAS  PubMed  Google Scholar 

  167. Dominguez-Sola, D. et al. Non-transcriptional control of DNA replication by c-Myc. Nature 448, 445–451 (2007).

    Article  CAS  PubMed  Google Scholar 

  168. Di Micco, R. et al. Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 444, 638–642 (2006).

    Article  CAS  PubMed  Google Scholar 

  169. Ohtsubo, M., Theodoras, A. M., Schumacher, J., Roberts, J. M. & Pagano, M. Human cyclin E, a nuclear protein essential for the G1-to-S phase transition. Mol. Cell. Biol. 15, 2612–2624 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Costantino, L. et al. Break-induced replication repair of damaged forks induces genomic duplications in human cells. Science 343, 88–91 (2014). This report shows that POLD3 is crucial for the survival of U2OS cells overexpressing cyclin E, suggesting that BIR is required to repair broken forks. Strikingly, POLD3-dependent repair generates gross chromosome rearrangements similar to those seen in human breast and ovarian cancers.

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  172. Aird, K. M. et al. Suppression of nucleotide metabolism underlies the establishment and maintenance of oncogene-induced senescence. Cell Rep. 3, 1252–1265 (2013).

    Article  CAS  PubMed  Google Scholar 

  173. Miron, K., Golan-Lev, T., Dvir, R., Ben-David, E. & Kerem, B. Oncogenes create a unique landscape of fragile sites. Nat. Commun. 6, 7094 (2015). The authors show that the overexpression of RAS or cyclin E in the same cells induces a set of fragile sites specific to each oncogene. These sites nest within large genes and colocalize with cancer breakpoints.

    Article  CAS  PubMed  Google Scholar 

  174. Lee, A. C. et al. Ras proteins induce senescence by altering the intracellular levels of reactive oxygen species. J. Biol. Chem. 274, 7936–7940 (1999).

    Article  CAS  PubMed  Google Scholar 

  175. 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).

    Article  CAS  PubMed  Google Scholar 

  176. Weyemi, U. et al. ROS-generating NADPH oxidase NOX4 is a critical mediator in oncogenic H-Ras-induced DNA damage and subsequent senescence. Oncogene 31, 1117–1129 (2012).

    Article  CAS  PubMed  Google Scholar 

  177. Ogrunc, M. et al. Oncogene-induced reactive oxygen species fuel hyperproliferation and DNA damage response activation. Cell Death Differ. 21, 998–1012 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Guo, Z., Kozlov, S., Lavin, M. F., Person, M. D. & Paull, T. T. ATM activation by oxidative stress. Science 330, 517–521 (2010).

    Article  CAS  PubMed  Google Scholar 

  179. Torres-Rosell, J. et al. Anaphase onset before complete DNA replication with intact checkpoint responses. Science 315, 1411–1415 (2007).

    Article  CAS  PubMed  Google Scholar 

  180. Naim, V. & Rosselli, F. The FANC pathway and mitosis: a replication legacy. Cell Cycle 8, 2907–2911 (2009).

    Article  CAS  PubMed  Google Scholar 

  181. Lukas, C. et al. 53BP1 nuclear bodies form around DNA lesions generated by mitotic transmission of chromosomes under replication stress. Nat. Cell Biol. 13, 243–253 (2011).

    Article  CAS  PubMed  Google Scholar 

  182. Lemmens, B., van Schendel, R. & Tijsterman, M. Mutagenic consequences of a single G-quadruplex demonstrate mitotic inheritance of DNA replication fork barriers. Nat. Commun. 6, 8909 (2015).

    Article  CAS  PubMed  Google Scholar 

  183. Maciejowski, J., Li, Y., Bosco, N., Campbell, P. J. & de Lange, T. Chromothripsis and kataegis induced by telomere crisis. Cell 163, 1641–1654 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Crosetto, N. et al. Nucleotide-resolution DNA double-strand break mapping by next-generation sequencing. Nat. Methods 10, 361–365 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Naim, V., Wilhelm, T., Debatisse, M. & Rosselli, F. ERCC1 and MUS81-EME1 promote sister chromatid separation by processing late replication intermediates at common fragile sites during mitosis. Nat. Cell Biol. 15, 1008–1015 (2013).

    Article  CAS  PubMed  Google Scholar 

  186. Ying, S. et al. MUS81 promotes common fragile site expression. Nat. Cell Biol. 15, 1001–1007 (2013). References 185 and 186 show that sequences remaining under-replicated in mitotic cells, notably at CFSs, are converted into DNA breaks by specialized nucleases such as MUS81–EME1.

    Article  CAS  PubMed  Google Scholar 

  187. Guervilly, J. H. et al. The SLX4 complex is a SUMO E3 ligase that impacts on replication stress outcome and genome stability. Mol. Cell 57, 123–137 (2015).

    Article  CAS  PubMed  Google Scholar 

  188. Bergoglio, V. et al. DNA synthesis by Pol eta promotes fragile site stability by preventing under-replicated DNA in mitosis. J. Cell Biol. 201, 395–408 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Minocherhomji, S. et al. Replication stress activates DNA repair synthesis in mitosis. Nature 528, 286–290 (2015).

    Article  CAS  PubMed  Google Scholar 

  190. Despras, E. et al. Rad18-dependent SUMOylation of human specialized DNA polymerase eta is required to prevent under-replicated DNA. Nat. Commun. 7, 13326 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Fernandez-Vidal, A. et al. A role for DNA polymerase theta in the timing of DNA replication. Nat. Commun. 5, 4285 (2014).

    Article  CAS  PubMed  Google Scholar 

  192. Hayano, M. et al. Rif1 is a global regulator of timing of replication origin firing in fission yeast. Genes Dev. 26, 137–150 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. Dave, A., Cooley, C., Garg, M. & Bianchi, A. Protein phosphatase 1 recruitment by Rif1 regulates DNA replication origin firing by counteracting DDK activity. Cell Rep. 7, 53–61 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Mattarocci, S. et al. Rif1 controls DNA replication timing in yeast through the PP1 phosphatase Glc7. Cell Rep. 7, 62–69 (2014).

    Article  CAS  PubMed  Google Scholar 

  195. Cornacchia, D. et al. Mouse Rif1 is a key regulator of the replication-timing programme in mammalian cells. EMBO J. 31, 3678–3690 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  196. Yamazaki, S. et al. Rif1 regulates the replication timing domains on the human genome. EMBO J. 31, 3667–3677 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Foti, R. et al. Nuclear architecture organized by Rif1 underpins the replication-timing program. Mol. Cell 61, 260–273 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Hassan-Zadeh, V. et al. USF binding sequences from the HS4 insulator element impose early replication timing on a vertebrate replicator. PLoS Biol. 10, e1001277 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Therizols, P. et al. Chromatin decondensation is sufficient to alter nuclear organization in embryonic stem cells. Science 346, 1238–1242 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. Hansel-Hertsch, R. et al. G-Quadruplex structures mark human regulatory chromatin. Nat. Genet. 48, 1267–1272 (2016).

    Article  CAS  PubMed  Google Scholar 

  201. Chambers, V. S. et al. High-throughput sequencing of DNA G-quadruplex structures in the human genome. Nat. Biotechnol. 33, 877–881 (2015).

    Article  PubMed  Google Scholar 

  202. Biffi, G., Tannahill, D., McCafferty, J. & Balasubramanian, S. Quantitative visualization of DNA G-quadruplex structures in human cells. Nat. Chem. 5, 182–186 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Henderson, A. et al. Detection of G-quadruplex DNA in mammalian cells. Nucleic Acids Res. 42, 860–869 (2014).

    Article  CAS  PubMed  Google Scholar 

  204. Castillo Bosch, P. et al. FANCJ promotes DNA synthesis through G-quadruplex structures. EMBO J. 33, 2521–2533 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  205. Ribeyre, C. et al. The yeast Pif1 helicase prevents genomic instability caused by G-quadruplex-forming CEB1 sequences in vivo. PLoS Genet. 5, e1000475 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  206. Piazza, A. et al. Genetic instability triggered by G-quadruplex interacting Phen-DC compounds in Saccharomyces cerevisiae. Nucleic Acids Res. 38, 4337–4348 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  207. Piazza, A. et al. Stimulation of gross chromosomal rearrangements by the human CEB1 and CEB25 minisatellites in Saccharomyces cerevisiae depends on G-quadruplexes or Cdc13. PLoS Genet. 8, e1003033 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Lopes, J. et al. G-Quadruplex-induced instability during leading-strand replication. EMBO J. 30, 4033–4046 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  209. Badie, S. et al. BRCA2 acts as a RAD51 loader to facilitate telomere replication and capping. Nat. Struct. Mol. Biol. 17, 1461–1469 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Zimmer, J. et al. Targeting BRCA1 and BRCA2 deficiencies with G-quadruplex-interacting compounds. Mol. Cell 61, 449–460 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Sarkies, P., Reams, C., Simpson, L. J. & Sale, J. E. Epigenetic instability due to defective replication of structured DNA. Mol. Cell 40, 703–713 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  212. Schiavone, D. et al. Determinants of G quadruplex-induced epigenetic instability in REV1-deficient cells. EMBO J. 33, 2507–2520 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  213. Papadopoulou, C., Guilbaud, G., Schiavone, D. & Sale, J. E. Nucleotide pool depletion induces G-quadruplex-dependent perturbation of gene expression. Cell Rep. 13, 2491–2503 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  214. Clynes, D. & Gibbons, R. J. ATRX and the replication of structured DNA. Curr. Opin. Genet. Dev. 23, 289–294 (2013).

    Article  CAS  PubMed  Google Scholar 

  215. Schiavone, D. et al. PrimPol is required for replicative tolerance of G quadruplexes in vertebrate cells. Mol. Cell 61, 161–169 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Pontarin, G. et al. Ribonucleotide reduction is a cytosolic process in mammalian cells independently of DNA damage. Proc. Natl Acad. Sci. USA 105, 17801–17806 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. Prem veer Reddy, G. & Pardee, A. B. Multienzyme complex for metabolic channeling in mammalian DNA replication. Proc. Natl Acad. Sci. USA 77, 3312–3316 (1980). This pioneer work shows that enzymes involved in dNTP and DNA synthesis physically interact, suggesting that precursors can be channelled to replication foci.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  218. Baril, E., Baril, B. & Laszlo, J. DNA polymerases in normal and regenerating rat liver and hepatomas: targets for chemotherapy. Adv. Enzyme Regul. 12, 355–372 (1974).

    Article  CAS  PubMed  Google Scholar 

  219. Wickremasinghe, R. G., Yaxley, J. C. & Hoffbrand, A. V. Gel filtration of a complex of DNA polymerase and DNA precursor-synthesizing enzymes from a human lymphoblastoid cell line. Biochim. Biophys. Acta 740, 243–248 (1983).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Taricani, L., Shanahan, F., Malinao, M. C., Beaumont, M. & Parry, D. A. Functional approach reveals a genetic and physical interaction between ribonucleotide reductase and CHK1 in mammalian cells. PLoS ONE 9, e111714 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  222. veer Reddy, G. P. & Pardee, A. B. Coupled ribonucleoside diphosphate reduction, channeling, and incorporation into DNA of mammalian cells. J. Biol. Chem. 257, 12526–12531 (1982).

    Article  CAS  PubMed  Google Scholar 

  223. Zhang, Y. W., Jones, T. L., Martin, S. E., Caplen, N. J. & Pommier, Y. Implication of checkpoint kinase-dependent up-regulation of ribonucleotide reductase R2 in DNA damage response. J. Biol. Chem. 284, 18085–18095 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  224. Sun, Y., Jiang, X., Chen, S., Fernandes, N. & Price, B. D. A role for the Tip60 histone acetyltransferase in the acetylation and activation of ATM. Proc. Natl Acad. Sci. USA 102, 13182–13187 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  225. Altmeyer, M. et al. Liquid demixing of intrinsically disordered proteins is seeded by poly(ADP-ribose). Nat. Commun. 6, 8088 (2015). This study shows that DNA breaks are the sites of the transient and reversible assembly of proteins with large intrinsically disordered domains, which creates dynamic nuclear pseudo-compartments.

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Ammazzalorso, F., Pirzio, L. M., Bignami, M., Franchitto, A. & Pichierri, P. ATR and ATM differently regulate WRN to prevent DSBs at stalled replication forks and promote replication fork recovery. EMBO J. 29, 3156–3169 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  228. 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  PubMed  PubMed Central  Google Scholar 

  229. 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  PubMed  PubMed Central  Google Scholar 

  230. Kais, Z. et al. FANCD2 maintains fork stability in BRCA1/2-deficient tumors and promotes alternative end-joining DNA repair. Cell Rep. 15, 2488–2499 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  231. Michl, J., Zimmer, J., Buffa, F. M., McDermott, U. & Tarsounas, M. FANCD2 limits replication stress and genome instability in cells lacking BRCA2. Nat. Struct. Mol. Biol. 23, 755–757 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  232. Ceccaldi, R., Sarangi, P. & D'Andrea, A. D. The Fanconi anaemia pathway: new players and new functions. Nat. Rev. Mol. Cell Biol. 17, 337–349 (2016).

    Article  CAS  PubMed  Google Scholar 

  233. Guervilly, J. H., Mace-Aime, G. & Rosselli, F. Loss of CHK1 function impedes DNA damage-induced FANCD2 monoubiquitination but normalizes the abnormal G2 arrest in Fanconi anemia. Hum. Mol. Genet. 17, 679–689 (2008).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

M.D.'s team is supported by the French Association pour la Recherche sur le Cancer (grant ARC- S1220130607073), Agence Nationale de la Recherche (grant ANR-13-BSV6-0008-01) and Institut National du Cancer (grant INCa-2013-103). A.N.'s team is supported by the grant ANR-14-CE35-0003-02.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michelle Debatisse.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

PowerPoint slides

Glossary

Replication stress

In this Review, defined as a condition that impedes the progression of replication forks strongly enough to prevent the maintenance of the normal replication rate through an increase in the density of the initiation event.

Replication origins

Chromosome regions from which replication starts, giving rise to bidirectional replication forks.

Topologically associated domains

(TADs). Sub-megabase-scale DNA domains comprising sequences that physically interact with each other more frequently than with regions located in adjacent domains. TADs represent spatially segregated genome regions.

Origin recognition complex

(ORC). A hetero-hexameric complex composed of six subunits (ORC1–6) that binds the DNA or the chromatin, which is a prerequisite for the subsequent building of functional replication origins.

DNA combing

A method stretching single DNA molecules on microscope coverslips, allowing their observation. In combination with in vivo labelling of newly synthesized DNA with analogues of DNA precursors, it allows the study of DNA replication.

Triplex DNA

A single-stranded DNA region bound to the major groove of the DNA duplex forming a three-stranded helix, which normally occurs at sequences with mirror symmetry.

Hairpins

DNA structures in which a strand folds on itself and forms intra-strand base pairing.

G-Quadruplexes

Four repeats of at least three guanines that can interact to form four-stranded DNA structures.

Nucleotide excision repair

(NER). Repair pathway that removes a broad range of lesions that distort the DNA double-helix structure. Processing relies on XPF–ERCC1 and XPG, two structure-dependent nucleases that remove a short oligonucleotide containing the lesions. The gap is then filled and repair completed.

Replication protein A

(RPA). This heterotrimeric complex specifically binds single-stranded DNA and protects it from degradation. RPA is a key factor in fork protection, ATR-CHK1 checkpoint activation, and DNA repair.

γH2AX

Phosphorylation of histone H2AX on serine 139 (γH2AX) by DNA damage response (DDR) apical kinases (ATM, ATR and DNAPK) constitutes a general marker of DNA damage and replication stress.

Synthetic lethality

Lethal phenotype that is associated with the co-inactivation of two proteins, whereas inactivation of either one of the proteins alone is viable. It is an important strategy for developing new cancer therapies.

Isolation of proteins on nascent DNA

(iPOND). A method that enables the purification of newly synthesized DNA and associated proteins upon immunoprecipitation of nucleotides analogues incorporated at the fork during a short pulse-labelling period.

Senescence

A phenomenon by which cells stop proliferating. It is usually induced by shortening of the telomeres and subsequent DNA damage response (DDR) activation.

MUS81–EME1/2

Structure-dependent nuclease that efficiently processes various DNA structures containing a junction between single-stranded DNA and double-stranded DNA, including replication forks. MUS81 is the catalytic subunit and EME1 and EME2 are two regulatory subunits that peak in G2/M and S phase cells, respectively.

MRE11

Protein belonging to a complex comprising MRE11, RAD50 and NBS1 (the MRN complex). The MRN complex is recruited to double-strand breaks (DSBs), where it has crucial roles in damage repair, and to replication forks upon replication stress. MRE11 has 3′–5′ exonuclease and single-stranded DNA (ssDNA) endonuclease activities that together contribute to DSB resection by producing the 3′ ssDNA overhang required in homology-directed repair and checkpoint activation.

p53 binding protein 1

(53BP1). Recruited to double-strand break (DSB) sites, where it promotes end-joining and inhibits homologous recombination repair. Large 53BP1 bodies are also formed upon replication stress, supposedly protecting DSBs from extensive degradation.

Break-induced replication

(BIR). A type of homologous recombination-driven replication that extends one-ended DNA double-strand breaks, which can form upon fork collapse. New DNA synthesis begins when the 3′ end of a break invades a double-strand template, forming a moving D-loop. Then, the second strand is synthesized.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Técher, H., Koundrioukoff, S., Nicolas, A. et al. The impact of replication stress on replication dynamics and DNA damage in vertebrate cells. Nat Rev Genet 18, 535–550 (2017). https://doi.org/10.1038/nrg.2017.46

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrg.2017.46

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