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.

Lagging-strand replication shapes the mutational landscape of the genome

Abstract

The origin of mutations is central to understanding evolution and of key relevance to health. Variation occurs non-randomly across the genome, and mechanisms for this remain to be defined. Here we report that the 5′ ends of Okazaki fragments have significantly increased levels of nucleotide substitution, indicating a replicative origin for such mutations. Using a novel method, emRiboSeq, we map the genome-wide contribution of polymerases, and show that despite Okazaki fragment processing, DNA synthesized by error-prone polymerase-α (Pol-α) is retained in vivo, comprising approximately 1.5% of the mature genome. We propose that DNA-binding proteins that rapidly re-associate post-replication act as partial barriers to Pol-δ-mediated displacement of Pol-α-synthesized DNA, resulting in incorporation of such Pol-α tracts and increased mutation rates at specific sites. We observe a mutational cost to chromatin and regulatory protein binding, resulting in mutation hotspots at regulatory elements, with signatures of this process detectable in both yeast and humans.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Increased substitution rates at OJs.
Figure 2: Frequent nucleotide substitutions at OF 5′ ends.
Figure 3: Mapping DNA synthesis in vivo using emRiboSeq.
Figure 4: Pol-α DNA synthesis contributes 1.5% of the mature genome.
Figure 5: OF mutational signatures are conserved in humans.

Accession codes

Primary accessions

Gene Expression Omnibus

Data deposits

Next-generation sequencing data submitted to the Gene Expression Omnibus (GEO) under accession number GSE64521; R and PERL code submitted to GitHub repository taylorLab/LaggingStrand (https://github.com/taylorLab/LaggingStrand).

References

  1. Kunkel, T. A. Evolving views of DNA replication (in)fidelity. Cold Spring Harb. Symp. Quant. Biol. 74, 91–101 (2009)

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Wolfe, K. H., Sharp, P. M. & Li, W. H. Mutation rates differ among regions of the mammalian genome. Nature 337, 283–285 (1989)

    ADS  CAS  PubMed  Google Scholar 

  3. Alexandrov, L. B. & Stratton, M. R. Mutational signatures: the patterns of somatic mutations hidden in cancer genomes. Curr. Opin. Genet. Dev. 24, 52–60 (2014)

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Ciccia, A. & Elledge, S. J. The DNA damage response: making it safe to play with knives. Mol. Cell 40, 179–204 (2010)

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Lindblad-Toh, K. et al. A high-resolution map of human evolutionary constraint using 29 mammals. Nature 478, 476–482 (2011)

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Pollard, K. S. et al. Forces shaping the fastest evolving regions in the human genome. PLoS Genet. 2, e168 (2006)

    PubMed  PubMed Central  Google Scholar 

  7. Prendergast, J. G. & Semple, C. A. Widespread signatures of recent selection linked to nucleosome positioning in the human lineage. Genome Res. 21, 1777–1787 (2011)

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Sasaki, S. et al. Chromatin-associated periodicity in genetic variation downstream of transcriptional start sites. Science 323, 401–404 (2009)

    ADS  CAS  PubMed  Google Scholar 

  9. Semple, C. A. & Taylor, M. S. Molecular biology. The structure of change. Science 323, 347–348 (2009)

    CAS  PubMed  Google Scholar 

  10. Warnecke, T., Batada, N. N. & Hurst, L. D. The impact of the nucleosome code on protein-coding sequence evolution in yeast. PLoS Genet. 4, e1000250 (2008)

    PubMed  PubMed Central  Google Scholar 

  11. Washietl, S., Machne, R. & Goldman, N. Evolutionary footprints of nucleosome positions in yeast. Trends Genet. 24, 583–587 (2008)

    CAS  PubMed  Google Scholar 

  12. Ying, H., Epps, J., Williams, R. & Huttley, G. Evidence that localized variation in primate sequence divergence arises from an influence of nucleosome placement on DNA repair. Mol. Biol. Evol. 27, 637–649 (2010)

    CAS  PubMed  Google Scholar 

  13. Johnston, L. H. & Nasmyth, K. A. Saccharomyces cerevisiae cell cycle mutant cdc9 is defective in DNA ligase. Nature 274, 891–893 (1978)

    ADS  CAS  PubMed  Google Scholar 

  14. Okazaki, R., Okazaki, T., Sakabe, K., Sugimoto, K. & Sugino, A. Mechanism of DNA chain growth. I. Possible discontinuity and unusual secondary structure of newly synthesized chains. Proc. Natl Acad. Sci. USA 59, 598–605 (1968)

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  15. Balakrishnan, L. & Bambara, R. A. Okazaki fragment metabolism. Cold Spring Harb. Perspect. Biol. 5, a010173 (2013)

    PubMed  PubMed Central  Google Scholar 

  16. Zheng, L. & Shen, B. Okazaki fragment maturation: nucleases take centre stage. J. Mol. Cell Biol. 3, 23–30 (2011)

    CAS  PubMed  PubMed Central  Google Scholar 

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

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  18. Stith, C. M., Sterling, J., Resnick, M. A., Gordenin, D. A. & Burgers, P. M. Flexibility of eukaryotic Okazaki fragment maturation through regulated strand displacement synthesis. J. Biol. Chem. 283, 34129–34140 (2008)

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Perera, R. L. et al. Mechanism for priming DNA synthesis by yeast DNA Polymerase alpha. Elife 2, e00482 (2013)

    PubMed  PubMed Central  Google Scholar 

  20. Walsh, E. & Eckert, K. A. Eukaryotic Replicative DNA Polymerases. Nucleic Acid Polymerases 30, 17–41 (2014)

    CAS  Google Scholar 

  21. Pavlov, Y. I. et al. Evidence that errors made by DNA polymerase α are corrected by DNA polymerase δ. Curr. Biol. 16, 202–207 (2006)

    CAS  PubMed  Google Scholar 

  22. Maga, G. et al. Okazaki fragment processing: modulation of the strand displacement activity of DNA polymerase δ by the concerted action of replication protein A, proliferating cell nuclear antigen, and flap endonuclease-1. Proc. Natl Acad. Sci. USA 98, 14298–14303 (2001)

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  23. Kunkel, T. A., Hamatake, R. K., Motto-Fox, J., Fitzgerald, M. P. & Sugino, A. Fidelity of DNA polymerase I and the DNA polymerase I-DNA primase complex from Saccharomyces cerevisiae. Mol. Cell. Biol. 9, 4447–4458 (1989)

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Lujan, S. A. et al. Heterogeneous polymerase fidelity and mismatch repair bias genome variation and composition. Genome Res. 24, 1751–1764 (2014)

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Nick McElhinny, S. A., Kissling, G. E. & Kunkel, T. A. Differential correction of lagging-strand replication errors made by DNA polymerases α and δ. Proc. Natl Acad. Sci. USA 107, 21070–21075 (2010)

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  26. Niimi, A. et al. Palm mutants in DNA polymerases α and η alter DNA replication fidelity and translesion activity. Mol. Cell. Biol. 24, 2734–2746 (2004)

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  Google Scholar 

  28. 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 

  29. Sparks, J. L. et al. RNase H2-initiated ribonucleotide excision repair. Mol. Cell 47, 980–986 (2012)

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Lujan, S. A., Williams, J. S., Clausen, A. R., Clark, A. B. & Kunkel, T. A. Ribonucleotides are signals for mismatch repair of leading-strand replication errors. Mol. Cell 50, 437–443 (2013)

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Lujan, S. A. et al. Mismatch repair balances leading and lagging strand DNA replication fidelity. PLoS Genet. 8, e1003016 (2012)

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Miyabe, I., Kunkel, T. A. & Carr, A. M. The major roles of DNA polymerases ε and δ at the eukaryotic replication fork are evolutionarily conserved. PLoS Genet. 7, e1002407 (2011)

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Nick McElhinny, S. A., Gordenin, D. A., Stith, C. M., Burgers, P. M. & Kunkel, T. A. Division of labor at the eukaryotic replication fork. Mol. Cell 30, 137–144 (2008)

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Larrea, A. A. et al. Genome-wide model for the normal eukaryotic DNA replication fork. Proc. Natl Acad. Sci. USA 107, 17674–17679 (2010)

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  35. Pursell, Z. F., Isoz, I., Lundstrom, E. B., Johansson, E. & Kunkel, T. A. Yeast DNA polymerase epsilon participates in leading-strand DNA replication. Science 317, 127–130 (2007)

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  36. Raghuraman, M. K. et al. Replication dynamics of the yeast genome. Science 294, 115–121 (2001)

    ADS  CAS  PubMed  Google Scholar 

  37. Nick McElhinny, S. A. et al. Abundant ribonucleotide incorporation into DNA by yeast replicative polymerases. Proc. Natl Acad. Sci. USA 107, 4949–4954 (2010)

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  38. Kao, H. I. & Bambara, R. A. The protein components and mechanism of eukaryotic Okazaki fragment maturation. Crit. Rev. Biochem. Mol. Biol. 38, 433–452 (2003)

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Boyle, A. P. et al. High-resolution genome-wide in vivo footprinting of diverse transcription factors in human cells. Genome Res. 21, 456–464 (2011)

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Cooper, G. M. et al. Distribution and intensity of constraint in mammalian genomic sequence. Genome Res. 15, 901–913 (2005)

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Schmidt, S. et al. Hypermutable non-synonymous sites are under stronger negative selection. PLoS Genet. 4, e1000281 (2008)

    PubMed  PubMed Central  Google Scholar 

  43. Vengrova, S. & Dalgaard, J. Z. The wild-type Schizosaccharomyces pombe mat1 imprint consists of two ribonucleotides. EMBO Rep. 7, 59–65 (2006)

    CAS  PubMed  Google Scholar 

  44. Ghodgaonkar, M. M. et al. Ribonucleotides misincorporated into DNA act as strand-discrimination signals in eukaryotic mismatch repair. Mol. Cell 50, 323–332 (2013)

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Liberti, S. E., Larrea, A. A. & Kunkel, T. A. Exonuclease 1 preferentially repairs mismatches generated by DNA polymerase α. DNA Repair 12, 92–96 (2013)

    CAS  PubMed  Google Scholar 

  46. Burgess, R. J. & Zhang, Z. Histone chaperones in nucleosome assembly and human disease. Nature Struct. Mol. Biol. 20, 14–22 (2013)

    CAS  Google Scholar 

  47. Villar, D., Flicek, P. & Odom, D. T. Evolution of transcription factor binding in metazoans - mechanisms and functional implications. Nature Rev. Genet. 15, 221–233 (2014)

    CAS  PubMed  Google Scholar 

  48. Clausen, A. R. et al. Tracking replication enzymology in vivo by genome-wide mapping of ribonucleotide incorporation. Nature Struct. Mol. Biol http://dx.doi.org/10.1038/nsmb.2957 (2015)

  49. Koh, K. D., Balachander, S., Hesselberth, J. R. & Storici, F. Ribose-seq: global mapping of ribonucleotides embedded in genomic DNA. Nature Methods http://dx.doi.org/10.1038/nmeth.3259 (2015)

  50. Daigaku, Y. et al. A global profile of replicative polymerase usage. Nature Struct. Mol. Biol. (in the press)

  51. Kuhn, R. M., Haussler, D. & Kent, W. J. The UCSC genome browser and associated tools. Brief. Bioinform. 14, 144–161 (2013)

    CAS  PubMed  Google Scholar 

  52. Derrien, T. et al. Fast computation and applications of genome mappability. PLoS ONE 7, e30377 (2012)

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  53. Eaton, M. L., Galani, K., Kang, S., Bell, S. P. & MacAlpine, D. M. Conserved nucleosome positioning defines replication origins. Genes Dev. 24, 748–753 (2010)

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Hesselberth, J. R. et al. Global mapping of protein-DNA interactions in vivo by digital genomic footprinting. Nature Methods 6, 283–289 (2009)

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Jiang, C. & Pugh, B. F. A compiled and systematic reference map of nucleosome positions across the Saccharomyces cerevisiae genome. Genome Biol. 10, R109 (2009)

    PubMed  PubMed Central  Google Scholar 

  56. Liti, G. et al. Population genomics of domestic and wild yeasts. Nature 458, 337–341 (2009)

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  57. Yang, Z. PAML 4: phylogenetic analysis by maximum likelihood. Mol. Biol. Evol. 24, 1586–1591 (2007)

    CAS  PubMed  Google Scholar 

  58. Rhee, H. S. & Pugh, B. F. Comprehensive genome-wide protein-DNA interactions detected at single-nucleotide resolution. Cell 147, 1408–1419 (2011)

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Korhonen, J., Martinmaki, P., Pizzi, C., Rastas, P. & Ukkonen, E. MOODS: fast search for position weight matrix matches in DNA sequences. Bioinformatics 25, 3181–3182 (2009)

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Mathelier, A. et al. JASPAR 2014: an extensively expanded and updated open-access database of transcription factor binding profiles. Nucleic Acids Res. 42, D142–D147 (2014)

    CAS  PubMed  Google Scholar 

  61. The ENCODE Project Consortium. An integrated encyclopedia of DNA elements in the human genome. Nature 489, 57–74 (2012)

  62. Thurman, R. E. et al. The accessible chromatin landscape of the human genome. Nature 489, 75–82 (2012)

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  63. Reijns, M. A. et al. The structure of the human RNase H2 complex defines key interaction interfaces relevant to enzyme function and human disease. J. Biol. Chem. 286, 10530–10539 (2011)

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank N. Hastie and P. Burgers for discussions, I. Adams, J. Caceres, T. Aitman and P. Heyn for comments on the manuscript, and A. Gallacher for technical assistance. We are indebted to J. Williams, A. Clausen and T. Kunkel for sharing yeast strains and unpublished data, and to S. Cerritelli and R. Crouch for RNH201 vectors. Funding: MRC Centenary Award to M.A.M.R.; MRC and Lister Institute of Preventive Medicine to A.P.J.; MRC and Medical Research Foundation to M.S.T.

Author information

Authors and Affiliations

Authors

Contributions

M.A.M.R. and J.D. performed experiments. H.K., S.M.d.P. and M.S.T. performed computational analysis. M.S.T., M.A.M.R. and A.P.J. designed the study, developed emRiboSeq and wrote the manuscript.

Corresponding authors

Correspondence to Andrew P. Jackson or Martin S. Taylor.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Extended data figures and tables

Extended Data Figure 1 Increased OJ and polymorphism rates correlate at binding sites of different nucleosome classes and at Rap1 binding sites.

af, OJ and polymorphism rates are strongly correlated for different classes of nucleosomes. Data presented as in Fig. 1a, for different sub-classes of S. cerevisiae nucleosomes, demonstrating that OJ and polymorphism rates co-vary in all cases. Transcription start site proximal nucleosomes (d) are probably subject to strong and asymmetrically distributed selective constraints, which is likely to explain the modestly reduced correlation for this subset. Such transcription start site proximal nucleosomes were excluded from analyses of other categories presented (b, c, e, f), except ‘all nucleosomes’ (a). g, OJ and polymorphism rates are correlated for the S. cerevisiae TF, Rap1. Data presented, as for Reb1 in Fig. 1b, show increased OJ and polymorphism rates around its binding site, with a dip corresponding to its central recognition sequence. hj, Increased polymorphism and OJ rates at Rap1 (h), nucleosome (i) and Reb1 (j) binding sites are not caused by biases in nucleotide content. Distributions calculated as for g, Fig. 1a and b, respectively, using a trinucleotide preserving genome shuffle. Pink shaded areas denote 95% confidence intervals for nucleotide substitution rates (100 shuffles). k, l, Polymorphism (red) and between-species (black) substitution rates are highly correlated for nucleosome (k) and Reb1 (l) binding sites. Best fit splines shown only. y axes scaled to demonstrate similar shape distribution. Values plotted as percentage relative to the mean rate for all data points (central 11 nucleotides excluded for calculation of mean in g, l).

Extended Data Figure 2 EmRiboSeq methodology and validation.

a, Schematic of emRiboSeq library preparation. rN, ribonucleotide. bd, Validation of strand-specific detection of enzymatically generated nicks through linker-ligation. Nb.BtsI nicking endonuclease cleaves the bottom strand of its recognition site releasing a 5′ fragment (cyan) with a free 3′-OH group after denaturation, to which the sequencing adaptor (pink) is ligated, allowing sequencing and mapping of this site to the genome (b). Nb.BtsI libraries have high reproducibility between Δrnh201 POL and Δrnh201 Pol-αpol1-L868M) strains after normalizing read counts to sequence tags per million (TPM). Bona fide Nb.BtsI sites were equally represented, at maximal frequency, in both libraries (c). Those with lower frequencies represented sites in close proximity to other Nb.BtsI sites, causing their partial loss during size selection. Additionally, Nb.BtsI-like sites were detected as the result of star activity. Libraries were also prepared using BciVI restriction enzyme digestion, that did not show such star activity (data not shown), allowing calculation of the site specificity for the method (>99.9%). Summed signal at Nb.BtsI sites shows >99.9% strand specificity (blue, correct strand; grey, opposite strand) and >99% single nucleotide resolution (d).

Extended Data Figure 3 Mapping replicative polymerase DNA synthesis using emRiboSeq.

a, Point mutations in replicative polymerases elevate ribonucleotide incorporation rates, permitting their contribution to genome synthesis to be tracked. Schematic of replication fork with polymerases and their ribonucleotide incorporation rates (refs 27, 30 and J. S. Williams, A. R. Clausen & T. A. Kunkel, personal communication) as indicated (POL denotes wild-type polymerases; asterisk denotes point mutants). Embedded ribonucleotides indicated by ‘R’; additional incorporation events due to polymerase mutations highlighted by shaded circles. b, c, Mapping of leading/lagging-strand synthesis by Pol-δ* and Pol-ε* yeast strain using emRiboSeq (as in Fig. 3) highlights both experimentally validated (pink dotted lines) and putative (grey dotted lines) replication origins. These often correspond to regions of early replicating DNA36 (c). d, Pol-αy axes show log2 of the strand ratio calculated in 2,001-nucleotide windows (bd).

Extended Data Figure 4 Quantification of in vivo ribonucleotide incorporation by replicative polymerases.

a, b, Representative alkaline gel electrophoresis of genomic DNA from yeast strains with mutant replicative DNA polymerases (a), with accompanying densitometry plots (b). Embedded ribonucleotides are detected by increased fragmentation of genomic DNA following alkaline treatment in an RNase H2-deficient (Δrnh201) background. Increased rates are seen with all three mutant polymerases (indicated by asterisk, as defined in Extended Data Fig. 3a), and are reduced in Pol-ε′ which contains the point mutation Met644Leu, a mutation that increases selectivity for dNTPs over rNTPs27. c, Quantification of average ribonucleotide incorporation in polymerase mutants from four independent experiments. DNA isolated from mid-log phase cultures; error bars denote s.e.m. Overall ribonucleotide content is the product of incorporation frequency and the total contribution of each polymerase, resulting in the total ribonucleotide content detected to be highest for Pol-εd, Most of the yeast genome exhibits directional asymmetry in replication (median 4:1 strand ratio). Count of genomic segments calculated for consecutive 2,001-nucleotide windows over the yeast genome based on reanalysis of OF sequencing data17 denoted as ‘Okazaki-seq’. The strand asymmetry ratio was calculated after re-orienting all regions such that the predominant lagging strand was the forward strand. eg, Genome-wide quantification of strand-specific incorporation of wild-type and mutant replicative DNA polymerases determined by emRiboSeq reflects their roles in leading- and lagging-strand replication. A close to linear correlation with Okazaki-seq strand ratios is observed. The strand ratio preference for lagging-strand ribonucleotide incorporation for independent libraries (including stationary phase libraries for POL and Pol-αf and g). f, g, Scatter plots illustrating the individual strand ratio data points for 2,001-nucleotide windows, for stationary phase POL (f) and Pol-αg) yeast. Pearson’s correlation = 0.49, P < 2.2 × 10−16 for POL (f); correlation = 0.75, P < 2.2 × 10−16 for Pol-αg).

Source data

Extended Data Figure 5 Pol-α-synthesized DNA retention is independent of RNase H2 processing of RNA primers.

a, b, The ribonucleotide content of genomic DNA is unchanged between Δrnh201 strains transformed with empty vector (−) or vector expressing Rnh201 separation-of-function mutant (sf), that retains the ability to cleave RNA:DNA hybrids, including RNA primers, but cannot cleave single embedded ribonucleotides39. In contrast, the same vector expressing wild-type Rnh201 (wt) fully rescues alkaline sensitivity of the DNA. As complementation with the separation-of-function mutant had no detectable effect on the ribonucleotide content seen in the Pol-α(Leu868Met) Δrnh201 strain, retention of Pol-α-synthesized DNA appears to be independent of a putative role for RNase H2 in RNA primer removal. Representative result shown for n = 3 independent experiments. c, Wild-type and mutant Rnh201 are expressed at equal levels, as shown by immuno-detection of the C-terminal FLAG tag. Loading control, actin.

Source data

Extended Data Figure 6 Elevated substitution rates are observed adjacent to many human TF binding sites.

ad, Nucleotide substitution rates (plotted as GERP scores) are elevated immediately adjacent to REST (a, b) and CTCF binding sites (c, d). Colour intensity shows quartiles of ChIP-seq peak height (pink to brown: lower to higher), reflecting strength of binding/occupancy. Stronger binding correlates with greater increases of proximal substitution rate in the ‘shoulder’ region (asterisk). Increased substitution rates are not a consequence of local sequence composition effects (b, d). Strongest binding quartile of sites (brown) is shown compared to a trinucleotide preserving shuffle (black) based on the flanking sequence (100–300 nucleotides from motif midpoint) of the same genomic locations. Brown dashed line and grey shading denote 95% confidence intervals. e, Substitution rates plotted as GERP scores for human TF binding sites identified in ChIP-seq data sets (in conjunction with binding site motif). Sites aligned (x = 0) on the midpoint of the TF binding site within the ChIP-seq peak (colours as for ad). Dashed black line shows y = 0, the genome wide expectation for neutral evolution.

Extended Data Figure 7 OJ and polymorphism rates are increased at yeast DNase I footprints.

a, b, DNase I footprint edges correspond, genome-wide, to increased OJ rates and locally elevated polymorphism rates in S. cerevisiae (a), a pattern that is maintained when footprints associated with Reb1 and Rap1 binding sites are excluded (b). Genome-wide DNase I footprints (n = 6,063) and excluding those within 50 nucleotides of a Reb1 or Rap1 binding site (n = 5,136) were aligned to their midpoint. c, d, Aligning DNase I footprints on their left edge rather than midpoint (to compensate for substantial heterogeneity in footprint size) demonstrates a distinct shoulder of elevated polymorphism rate at the aligned edge (c), with a significant elevation compared to nearby sequence upstream from the footprint (d). DNase I footprints from a were aligned to their left edge (x = 0) with corresponding polymorphism rates shown (c). The increased polymorphism rate cannot be explained by local sequence compositional distortions (d). Nucleotide substitution rates in the 11 nucleotides centred on the DNase footprint edge (pink line), and another 11 nucleotides encompassing positions −35 to −25 relative to the footprint edge (green line) were quantified. Darker pink and green filled circles denote the mean of observed substitution rates and lighter shades denote the mean for the same sites after trinucleotide preserving genomic shuffles. Error bars denote s.d.; statistics by Mann–Whitney test. e, Model shows that correlation of increased nucleotide substitution and OJ rates are consistent with increased mutation frequency across heterogeneous DNase I footprints. Polymorphism is reduced at sequence-specific binding sites within the footprints, owing to functional constraint. Therefore, the effect of OF-related mutagenesis in these regions is most sensitively detected in the region immediately adjacent to the binding site (left of vertical dashed blue line, representing footprints aligned to their left edge). This ‘shoulder’ of increased nucleotide substitutions represents sites with increased, OJ-associated mutation is followed by a region of depressed substitution rates, owing to selective effects of the functional binding sites within the footprints (to the right of the dashed blue line). Signals further to the right are not interpretable given the heterogeneity in DNase I footprint sizes. Given strong selection at TF and DNase I footprint sites, this ‘shoulder’ of elevated nucleotide substitutions could represent a measure for the local mutation rate for such regions, analogous to that measured by the fourfold degenerate sites in protein coding sequence.

Extended Data Figure 8 Model to show Pol-α DNA tract retention downstream of protein binding sites.

a, OF priming occurs stochastically, with the 5′ end of each OF initially synthesized by Pol-α and the remainder of the OF synthesized by Pol-δ. b, c, OF processing: when Pol-δ encounters the previously synthesized OF, Pol-δ continues to synthesize DNA displacing the 5′ end of the downstream OF, which is removed by nucleases to result in mature OFs which are then ligated. The OJs of such mature OFs before ligation were detected previously17 after depletion of temperature-sensitive DNA ligase I. They demonstrated that if a protein barrier is encountered (grey circle), Pol-δ progression is impaired, leading to reduced removal of the downstream OF (b). Given that 1.5% of the mature genome is synthesized by Pol-α, a proportion of lagging strands will retain Pol-α-synthesized DNA (red). When Pol-δ progression is impaired by protein binding, this will lead to an increased fraction of fragments containing Pol-α-synthesized DNA downstream of such sites (c).

Supplementary information

Supplementary Information

This file contains Supplementary Tables 1-3 and additional references. (PDF 175 kb)

PowerPoint slides

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Reijns, M., Kemp, H., Ding, J. et al. Lagging-strand replication shapes the mutational landscape of the genome. Nature 518, 502–506 (2015). https://doi.org/10.1038/nature14183

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature14183

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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