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.

  • Article
  • Published:

Transcription shapes DNA replication initiation and termination in human cells

Abstract

Although DNA replication is a fundamental aspect of biology, it is not known what determines where DNA replication starts and stops in the human genome. We directly identified and quantitatively compared sites of replication initiation and termination in untransformed human cells. We found that replication preferentially initiates at the transcription start site of genes occupied by high levels of RNA polymerase II, and terminates at their polyadenylation sites, thereby ensuring global co-directionality of transcription and replication, particularly at gene 5′ ends. During replication stress, replication initiation is stimulated downstream of genes and termination is redistributed to gene bodies; this globally reorients replication relative to transcription around gene 3′ ends. These data suggest that replication initiation and termination are coupled to transcription in human cells, and propose a model for the impact of replication stress on genome integrity.

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

Fig. 1: DNA replication initiates preferentially at transcription start sites.
Fig. 2: Total RNA polymerase occupancy of a gene predicts replication origin firing efficiency at its TSS.
Fig. 3: Replication initiates at enhancers, largely independent of enhancer transcription.
Fig. 4: Global modulation of origin activity under conditions that increase or suppress dormant origin firing.
Fig. 5: Widespread, R-loop-independent replication-fork termination occurs at the 3′ ends of transcribed genes under unperturbed conditions.
Fig. 6: Redistribution of replication termination to gene bodies under replication stress.
Fig. 7: Model.

Similar content being viewed by others

Data availability

Data, including raw sequencing reads and tables used to generate source data for graphs in Figs. 16, are publicly available under accession number GSE114017. Custom scripts are available upon request from the corresponding authors.

References

  1. Stinchcomb, D. T., Struhl, K. & Davis, R. W. Isolation and characterisation of a yeast chromosomal replicator. Nature 282, 39–43 (1979).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  4. Besnard, E. et al. Unraveling cell type-specific and reprogrammable human replication origin signatures associated with G-quadruplex consensus motifs. Nat. Struct. Mol. Biol. 19, 837–844 (2012).

    Article  CAS  Google Scholar 

  5. Dellino, G. I. et al. Genome-wide mapping of human DNA-replication origins: levels of transcription at ORC1 sites regulate origin selection and replication timing. Genome Res. 23, 1–11 (2013).

    Article  CAS  Google Scholar 

  6. Langley, A. R., Gräf, S., Smith, J. C. & Krude, T. Genome-wide identification and characterisation of human DNA replication origins by initiation site sequencing (ini-seq). Nucleic Acids Res. 44, 10230–10247 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Petryk, N. et al. Replication landscape of the human genome. Nat. Commun. 7, 10208 (2016).

    Article  CAS  Google Scholar 

  8. Donovan, S., Harwood, J., Drury, L. S. & Diffley, J. F. Cdc6p-dependent loading of Mcm proteins onto pre-replicative chromatin in budding yeast. Proc. Natl Acad. Sci. USA 94, 5611–5616 (1997).

    Article  CAS  Google Scholar 

  9. Edwards, M. C. et al. MCM2-7 complexes bind chromatin in a distributed pattern surrounding the origin recognition complex in Xenopus egg extracts. J. Biol. Chem. 277, 33049–33057 (2002).

    Article  CAS  Google Scholar 

  10. 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  Google Scholar 

  11. Rocha, E. P. C. Gene essentiality determines chromosome organisation in bacteria. Nucleic Acids Res. 31, 6570–6577 (2003).

    Article  CAS  Google Scholar 

  12. Osmundson, J. S., Kumar, J., Yeung, R. & Smith, D. J. Pif1-family helicases cooperatively suppress widespread replication-fork arrest at tRNA genes. Nat. Struct. Mol. Biol. 24, 162–170 (2017).

    Article  CAS  Google Scholar 

  13. Pourkarimi, E., Bellush, J. M. & Whitehouse, I. Spatiotemporal coupling and decoupling of gene transcription with DNA replication origins during embryogenesis in C. elegans. eLife 5, e21728 (2016).

  14. Hamperl, S., Bocek, M. J., Saldivar, J. C., Swigut, T. & Cimprich, K. A. Transcription-replication conflict orientation modulates r-loop levels and activates distinct dna damage responses. Cell 170, 774–786.e19 (2017).

    Article  CAS  Google Scholar 

  15. Tran, P. L. T. et al. PIF1 family DNA helicases suppress R-loop mediated genome instability at tRNA genes. Nat. Commun. 8, 15025 (2017).

    Article  Google Scholar 

  16. McGuffee, S. R., Smith, D. J. & Whitehouse, I. Quantitative, genome-wide analysis of eukaryotic replication initiation and termination. Mol. Cell 50, 123–135 (2013).

    Article  CAS  Google Scholar 

  17. Tubbs, A. et al. Dual roles of poly(dA:dT) tracts in replication initiation and fork collapse. Cell 174, 1127–1142.e19 (2018).

    Article  CAS  Google Scholar 

  18. Smith, D. J. & Whitehouse, I. Intrinsic coupling of lagging-strand synthesis to chromatin assembly. Nature 483, 434–438 (2012).

    Article  CAS  Google Scholar 

  19. Harenza, J. L. et al. Transcriptomic profiling of 39 commonly-used neuroblastoma cell lines. Sci. Data 4, 170033 (2017).

    Article  CAS  Google Scholar 

  20. Macheret, M. & Halazonetis, T. D. Intragenic origins due to short G1 phases underlie oncogene-induced DNA replication stress. Nature 555, 112–116 (2018).

    Article  CAS  Google Scholar 

  21. Almeida, R. et al. Chromatin conformation regulates the coordination between DNA replication and transcription. Nat. Commun. 9, 1590 (2018).

    Article  Google Scholar 

  22. Sanchez, G. J. et al. Genome-wide dose-dependent inhibition of histone deacetylases studies reveal their roles in enhancer remodeling and suppression of oncogenic super-enhancers. Nucleic Acids Res. 46, 1756–1776 (2018).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  24. Karnani, N. & Dutta, A. The effect of the intra-S-phase checkpoint on origins of replication in human cells. Genes Dev. 25, 621–633 (2011).

    Article  CAS  Google Scholar 

  25. 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  Google Scholar 

  26. Michl, J., Zimmer, J. & Tarsounas, M. Interplay between Fanconi anemia and homologous recombination pathways in genome integrity. EMBO J. 35, 909–923 (2016).

    Article  CAS  Google Scholar 

  27. 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  Google Scholar 

  28. Proudfoot, N. J. Transcriptional termination in mammals: stopping the RNA polymerase II juggernaut. Science 352, aad9926 (2016).

    Article  Google Scholar 

  29. 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  Google Scholar 

  30. Merrikh, H. Spatial and temporal control of evolution through replication-transcription conflicts. Trends Microbiol. 25, 515–521 (2017).

    Article  CAS  Google Scholar 

  31. Paul, S., Million-Weaver, S., Chattopadhyay, S., Sokurenko, E. & Merrikh, H. Accelerated gene evolution through replication-transcription conflicts. Nature 495, 512–515 (2013).

    Article  CAS  Google Scholar 

  32. Srivatsan, A., Tehranchi, A., MacAlpine, D. M. & Wang, J. D. Co-orientation of replication and transcription preserves genome integrity. PLoS Genet. 6, e1000810 (2010).

    Article  Google Scholar 

  33. Hansen, R. S. et al. Sequencing newly replicated DNA reveals widespread plasticity in human replication timing. Proc. Natl Acad. Sci. USA 107, 139–144 (2009).

    Article  Google Scholar 

  34. Anderson, J. D. & Widom, J. Poly(dA-dT) promoter elements increase the equilibrium accessibility of nucleosomal DNA target sites. Mol. Cell. Biol. 21, 3830–3839 (2001).

    Article  CAS  Google Scholar 

  35. Vashee, S. et al. Sequence-independent DNA binding and replication initiation by the human origin recognition complex. Genes Dev. 17, 1894–1908 (2003).

    Article  CAS  Google Scholar 

  36. Glover-Cutter, K., Kim, S., Espinosa, J. & Bentley, D. L. RNA polymerase II pauses and associates with pre-mRNA processing factors at both ends of genes. Nat. Struct. Mol. Biol. 15, 71–78 (2008).

    Article  CAS  Google Scholar 

  37. Mirkin, E. V., Castro Roa, D., Nudler, E. & Mirkin, S. M. Transcription regulatory elements are punctuation marks for DNA replication. Proc. Natl Acad. Sci. USA 103, 7276–7281 (2006).

    Article  CAS  Google Scholar 

  38. Lang, K. S. et al. Replication-transcription conflicts generate R-loops that orchestrate bacterial stress survival and pathogenesis. Cell 170, 787–799.e18 (2017).

    Article  CAS  Google Scholar 

  39. 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).

    Article  CAS  Google Scholar 

  40. Stirling, P. C. et al. R-loop-mediated genome instability in mRNA cleavage and polyadenylation mutants. Genes Dev. 26, 163–175 (2012).

    Article  CAS  Google Scholar 

  41. Gros, J., Devbhandari, S. & Remus, D. Origin plasticity during budding yeast DNA replication in vitro. EMBO J. 33, 621–636 (2014).

    Article  CAS  Google Scholar 

  42. 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  Google Scholar 

  43. Douglas, M. E., Ali, F. A., Costa, A. & Diffley, J. F. X. The mechanism of eukaryotic CMG helicase activation. Nature 555, 265–268 (2018).

    Article  CAS  Google Scholar 

  44. 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  Google Scholar 

  45. Klein, K. et al. Genome-wide identification of early-firing human replication origins by optical replication mapping. Preprint at https://www.biorxiv.org/content/early/2017/11/06/214841 (2017).

Download references

Acknowledgements

We thank the NYU Genome Technology Center for assistance with TapeStation and sequencing. We thank D. Remus, I. Whitehouse, E. Mazzoni, and H. Klein for helpful discussions, and G Sanchez for sharing RPE-1 enhancer data. Y.-H.C. was funded in part by the Molecular Oncology and Immunology NCI training program through NYU School of Medicine (no. 5T32CA009161-40). Work in T.T.H.'s laboratory is supported by grants from the NIH (ES025166), V Foundation BRCA Research and Basser Innovation Award. Work in D.J.S.'s laboratory is supported by grants from the NIH (nos. GM127336, GM114340) and the Searle Scholars Program.

Author information

Authors and Affiliations

Authors

Contributions

Y.-H. C., M.K., and P.T. performed the experiments, S.K. and D.J.S. analyzed the data and D.F., T.T.H., and D.J.S. conceived and supervised the study. All of the authors interpreted the data. D.J.S. wrote the manuscript with input from all authors.

Corresponding authors

Correspondence to David Fenyö, Tony T. Huang or Duncan J. Smith.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Integrated supplementary information

Supplementary Fig 1 Reproducibility of TSS data across replicate datasets.

a, Percentage of Okazaki fragments (OFs) mapping to the Crick strand across a region ±50 kb from random genomic loci. b, Percentage of OFs mapping to the Crick strand across a ±50-kb window around the TSS of Watson (W) or Crick (C) genes. Data were analyzed as in Fig. 1e for two replicate datasets. c, Percentage of replication forks moving left to right around TSS binned by transcriptional volume (FPKM from 19 × gene length). Data were analyzed as in Fig. 2g, for two replicate datasets. d, Percentage of replication forks moving left to right around TSS binned by transcriptional volume, for cells treated with siRNAs against FANCD2 (green), FANCI (blue), or mock-treated (black), grown in 0.2 mM hydroxyurea for 4 h before OF collection. Data were analyzed as in Fig. 4d, using the second replicate datasets for each knockdown condition.

Supplementary Fig 2 The effect of gene length on TSS-proximal origin firing efficiency is not solely a result of passive replication.

a, Percentage of replication forks moving left to right around the TSS of actively transcribed genes (FPKM > median), where the TSS of the most proximal upstream gene is under (black) or over (green) 50 kb from the TSS being analyzed. b, Percentage of replication forks moving left to right around the TSS of actively transcribed genes (FPKM > median), where the TTS of the most proximal upstream gene is under (black) or over (red) 50 kb from the TSS being analyzed. c, Percentage of replication forks moving left to right around the TSS of actively transcribed genes (FPKM > median), where the TSS of the most proximal downstream gene is under (black) or over (green) 50 kb from the TSS being analyzed. d, Percentage of replication forks moving left to right around the TSS of actively transcribed genes (FPKM > median), where the most proximal downstream TTS is under (black) or over (red) 50 kb from the TSS being analyzed. Note that this TSS–TTS distance is equivalent to the length of the gene.

Supplementary Fig 3 Replication initiation is most efficient at high-volume TSS in HeLa and GM06990 cells.

Data in af were analyzed using Ok-seq data from HeLa cells (7) and HeLa RNA-seq data. a, Percentage of HeLa OFs mapping to the Crick strand (indicating rightward-moving replication forks) across a ±50-kb window around annotated TSS. Data were analyzed as in Fig. 1d. b, Percentage of OFs mapping to the Crick strand across a ±50-kb window around the TSS of Watson (W) or Crick (C) genes. Data were analyzed as in Fig. 1e. c, Percentage of replication forks moving from left to right around the TSS of all genes, oriented such that transcription runs from left to right. Data were analyzed as in Fig. 1f. d, Replication initiation frequency, calculated as the first derivative of Okazaki fragment strand bias as a function of position, across a ±50-kb window around the TSS. Data were analyzed as in Fig. 1g. e, Percentage of replication forks moving left to right around TSS binned by RNA-seq read depth quartile. Data were analyzed as in Fig. 2a. f, Percentage of replication forks moving left to right around TSS binned by gene length. Data were analyzed as in Fig. 2e. gl, Data were analyzed as in af, using Ok-seq data from GM06990 cells (7) and GM06990 RNA-seq data.

Supplementary Fig 4 Comparison of OF strand bias around enhancers and equivalently selected random sites at various distances from TSS.

Percentage of OFs mapping to the Crick strand around enhancer midpoints, for enhancers or random sites within the indicated distance of annotated TSS. Enhancers are binned according to transcription level (above or below median). Data were analyzed as in Fig. 3c.

Supplementary Fig 5 The effect of FANCI knockdown is not related to gene length or increased replication termination.

Percentage of replication forks moving left to right around TSS binned by transcriptional volume (FPKM from Harenza et al. × gene length) for cells treated with siRNAs against FANCI (blue) or mock-treated (black), grown in 0.2 mM hydroxyurea for 4 h before OF collection. The TSS is denoted by a gray dotted line: lower and upper bounds for gene length in each quartile are denoted by dotted and dashed black lines, respectively.

Supplementary Fig 6 Reproducibility of TTS data across replicate datasets.

a, Percentage of replication forks moving left to right around transcription termination sites (TTS) binned by RNA-seq read depth quartile. Data were analyzed as in Fig. 5a, for two replicate datasets. b, Percentage of replication forks moving left to right around (TTS) binned by RNA-seq read depth quartile, from cells grown in 0.2 mM HU for 4 h before OF collection. Data were analyzed as in Fig. 6a, for two replicate datasets.

Supplementary Fig 7 Transcription-dependent, R-loop-independent replication termination at TTS in HeLa and GM06990 cells.

Data in ad were analyzed using Ok-seq data from HeLa cells (7) and HeLa RNA-seq data. DRIP-seq data are from HeLa cells (14). a, Percentage of replication forks moving left to right around transcription termination sites (TTS) binned by RNA-seq read depth quartile. Data were analyzed as in Fig. 5a. b, Replication initiation frequency, calculated as the first derivative of Okazaki fragment strand bias, around TTS binned by RNA-seq read density in the gene body. Data were analyzed as in Fig. 5c. c, Percentage of replication forks moving left to right around TTS of actively transcribed (FPKM > median) high-DRIP versus low-DRIP genes. Data were analyzed as in Fig. 5g. d, Replication initiation frequency, calculated as the first derivative of Okazaki fragment strand bias, around TTS of actively transcribed (FPKM > median) high-DRIP versus low-DRIP genes. Data were analyzed as in Fig. 5h. eh, Data were analyzed as in ad using Ok-seq data from GM06990 cells (7) and GM06990 RNA-seq data.

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, YH., Keegan, S., Kahli, M. et al. Transcription shapes DNA replication initiation and termination in human cells. Nat Struct Mol Biol 26, 67–77 (2019). https://doi.org/10.1038/s41594-018-0171-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-018-0171-0

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing