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:

Evidence that processing of ribonucleotides in DNA by topoisomerase 1 is leading-strand specific

Abstract

Ribonucleotides incorporated during DNA replication are removed by RNase H2–dependent ribonucleotide excision repair (RER). In RER-defective yeast, topoisomerase 1 (Top1) incises DNA at unrepaired ribonucleotides, initiating their removal, but this is accompanied by RNA-DNA–damage phenotypes. Here we show that these phenotypes are incurred by a high level of ribonucleotides incorporated by a leading strand–replicase variant, DNA polymerase (Pol) ɛ, but not by orthologous variants of the lagging-strand replicases, Pols α or δ. Moreover, loss of both RNases H1 and H2 is lethal in combination with increased ribonucleotide incorporation by Pol ɛ but not by Pols α or δ. Several explanations for this asymmetry are considered, including the idea that Top1 incision at ribonucleotides relieves torsional stress in the nascent leading strand but not in the nascent lagging strand, in which preexisting nicks prevent the accumulation of superhelical tension.

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: Ribonucleotide incorporation in vitro by variants of Pols α and δ.
Figure 2: Strand-specific probing for ribonucleotides in nascent lagging-strand genomic DNA.
Figure 3: Lack of Top1-initiated ribonucleotide removal in pol1-L868M rnh201Δ and pol3-L612M rnh201Δ strains.
Figure 4: RNase H2 is dispensable for maintaining genome integrity in strains with increased capacity to incorporate ribonucleotides into lagging-strand DNA.
Figure 5: A model depicting three possibilities for strand-specific consequences of unrepaired ribonucleotides in the genomes of RER-defective yeast strains.

Similar content being viewed by others

References

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Wang, J. et al. Crystal structure of a pol alpha family replication DNA polymerase from bacteriophage RB69. Cell 89, 1087–1099 (1997).

    Article  CAS  PubMed  Google Scholar 

  5. Pavlov, Y.I., Shcherbakova, P.V. & Kunkel, T.A. In vivo consequences of putative active site mutations in yeast DNA polymerases alpha, epsilon, delta, and zeta. Genetics 159, 47–64 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. 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 

  7. Eder, P.S., Walder, R.Y. & Walder, J.A. Substrate specificity of human RNase H1 and its role in excision repair of ribose residues misincorporated in DNA. Biochimie 75, 123–126 (1993).

    Article  CAS  PubMed  Google Scholar 

  8. Rydberg, B. & Game, J. Excision of misincorporated ribonucleotides in DNA by RNase H (type 2) and FEN-1 in cell-free extracts. Proc. Natl. Acad. Sci. USA 99, 16654–16659 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Williams, J.S. et al. Topoisomerase 1-mediated removal of ribonucleotides from nascent leading-strand DNA. Mol. Cell 49, 1010–1015 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  16. Dalgaard, J.Z. Causes and consequences of ribonucleotide incorporation into nuclear DNA. Trends Genet. 28, 592–597 (2012).

    Article  CAS  PubMed  Google Scholar 

  17. Williams, J.S. & Kunkel, T.A. Ribonucleotides in DNA: origins, repair and consequences. DNA Repair (Amst.) 19, 27–37 (2014).

    Article  CAS  PubMed Central  Google Scholar 

  18. Caldecott, K.W. Ribose: an internal threat to DNA. Science 343, 260–261 (2014).

    Article  CAS  PubMed  Google Scholar 

  19. Tumbale, P., Williams, J.S., Schellenberg, M.J., Kunkel, T.A. & Williams, R.S. Aprataxin resolves adenylated RNA-DNA junctions to maintain genome integrity. Nature 506, 111–115 (2014).

    Article  CAS  PubMed  Google Scholar 

  20. Sekiguchi, J. & Shuman, S. Site-specific ribonuclease activity of eukaryotic DNA topoisomerase I. Mol. Cell 1, 89–97 (1997).

    Article  CAS  PubMed  Google Scholar 

  21. Kim, N. et al. Mutagenic processing of ribonucleotides in DNA by yeast topoisomerase I. Science 332, 1561–1564 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Clark, A.B., Lujan, S.A., Kissling, G.E. & Kunkel, T.A. Mismatch repair-independent tandem repeat sequence instability resulting from ribonucleotide incorporation by DNA polymerase epsilon. DNA Repair (Amst.) 10, 476–482 (2011).

    Article  CAS  PubMed Central  Google Scholar 

  23. Cho, J.E., Kim, N., Li, Y.C. & Jinks-Robertson, S. Two distinct mechanisms of Topoisomerase 1-dependent mutagenesis in yeast. DNA Repair (Amst.) 12, 205–211 (2013).

    Article  CAS  Google Scholar 

  24. Lazzaro, F. et al. RNase H and postreplication repair protect cells from ribonucleotides incorporated in DNA. Mol. Cell 45, 99–110 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Cerritelli, S.M. & Crouch, R.J. Ribonuclease H: the enzymes in eukaryotes. FEBS J. 276, 1494–1505 (2009).

    Article  CAS  PubMed  Google Scholar 

  26. 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 

  27. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Clausen, A.R. et al. Tracking replication enzymology in vivo by genome-wide mapping of ribonucleotide incorporation. Nat. Struct. Mol. Biol. 22, 185–191 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Burgers, P.M. Polymerase dynamics at the eukaryotic DNA replication fork. J. Biol. Chem. 284, 4041–4045 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Marteijn, J.A., Lans, H., Vermeulen, W. & Hoeijmakers, J.H. Understanding nucleotide excision repair and its roles in cancer and ageing. Nat. Rev. Mol. Cell Biol. 15, 465–481 (2014).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Garg, P., Stith, C.M., Sabouri, N., Johansson, E. & Burgers, P.M. Idling by DNA polymerase delta maintains a ligatable nick during lagging-strand DNA replication. Genes Dev. 18, 2764–2773 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Aguilera, A. & Klein, H.L. Genetic control of intrachromosomal recombination in Saccharomyces cerevisiae. I. Isolation and genetic characterization of hyper-recombination mutations. Genetics 119, 779–790 (1988).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Potenski, C.J., Niu, H., Sung, P. & Klein, H.L. Avoidance of ribonucleotide-induced mutations by RNase H2 and Srs2-Exo1 mechanisms. Nature 511, 251–254 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Williams, J.S. et al. Proofreading of ribonucleotides inserted into DNA by yeast DNA polymerase epsilon. DNA Repair (Amst.) 11, 649–656 (2012).

    Article  CAS  PubMed Central  Google Scholar 

  40. Wang, J.C. Cellular roles of DNA topoisomerases: a molecular perspective. Nat. Rev. Mol. Cell Biol. 3, 430–440 (2002).

    Article  CAS  PubMed  Google Scholar 

  41. Yu, H. & Droge, P. Replication-induced supercoiling: a neglected DNA transaction regulator? Trends Biochem. Sci. 39, 219–220 (2014).

    Article  CAS  PubMed  Google Scholar 

  42. Nilsen, H. et al. Uracil-DNA glycosylase (UNG)-deficient mice reveal a primary role of the enzyme during DNA replication. Mol. Cell 5, 1059–1065 (2000).

    Article  CAS  PubMed  Google Scholar 

  43. Bermejo, R. et al. Top1- and Top2-mediated topological transitions at replication forks ensure fork progression and stability and prevent DNA damage checkpoint activation. Genes Dev. 21, 1921–1936 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Gambus, A. et al. GINS maintains association of Cdc45 with MCM in replisome progression complexes at eukaryotic DNA replication forks. Nat. Cell Biol. 8, 358–366 (2006).

    Article  CAS  PubMed  Google Scholar 

  45. Georgescu, R.E. et al. Mechanism of asymmetric polymerase assembly at the eukaryotic replication fork. Nat. Struct. Mol. Biol. 21, 664–670 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Burgers, P.M. & Gerik, K.J. Structure and processivity of two forms of Saccharomyces cerevisiae DNA polymerase delta. J. Biol. Chem. 273, 19756–19762 (1998).

    Article  CAS  PubMed  Google Scholar 

  47. Shcherbakova, P.V. & Kunkel, T.A. Mutator phenotypes conferred by MLH1 overexpression and by heterozygosity for mlh1 mutations. Mol. Cell. Biol. 19, 3177–3183 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Pavlov, Y.I., Newlon, C.S. & Kunkel, T.A. Yeast origins establish a strand bias for replicational mutagenesis. Mol. Cell 10, 207–213 (2002).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank K. Bebenek, C. Orebaugh and S. Williams for helpful comments on the manuscript and all members of the Kunkel laboratory for thoughtful discussions. We acknowledge the US National Institute of Environmental Health Sciences (NIEHS) Molecular Genetics Core Facility for sequence analysis of 5-FOA–resistant mutants and the NIEHS Flow Cytometry Center for fluorescence-activated cell-sorting analysis. This work was supported by project Z01 ES065070 to T.A.K. from the Division of Intramural Research of the US National Institutes of Health (NIH), NIEHS, by the Swedish Cancer Society to A.C. and by US NIH grant GM032431 to P.M.B.

Author information

Authors and Affiliations

Authors

Contributions

J.S.W., A.R.C., S.A.L., L.M., A.B.C. and A.C. designed and performed experiments. P.M.B. provided reagents. All authors were involved in data analysis. J.S.W. and T.A.K. wrote the manuscript, with input from all authors.

Corresponding author

Correspondence to Thomas A Kunkel.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Probing for ribonucleotides in nascent leading-strand DNA.

(a) The annealing locations of radiolabeled leading strand-specific probes on the URA3 reporter in OR1 or OR2. (b) Detection of alkali-sensitive sites in nascent leading strand DNA synthesized by Pol ɛ was performed as described13. Smaller DNA fragments observed for the α-LM rnh201Δ and δ-LM rnh201Δ mutants when using probes that anneal to the nascent leading strand (lanes 4, 6, 12 and 14) may be related to the close proximity of URA3 to ARS306 (1.6 kb). These alkali-sensitive sites may arise during ribonucleotide incorporation by Pols α or δ into the nascent lagging strand during bidirectional synthesis proceeding from this origin in the opposite direction (to the left of the origin in panel a). In addition, these small fragments that hybridize to the ‘leading strand’ probe may be generated by synthesis performed by L868M Pol α or L612M Pol δ as they replicate from the adjacent ARS307 origin. The same explanation applies for small DNA fragments observed for the ɛ-MG rnh201Δ mutant in Figure 2 (main text) when using probes that anneal to the nascent lagging strand (lanes 8 and 16). (c) The average fraction of alkali-sensitive fragments along the membrane was determined by quantifying the radioactive signal using data from in panel b (for both Probes A and B) from two independent experiments.

Supplementary Figure 2 Workflow for calculating fractional replication-strand bias from HydEn-seq end counts without background subtraction or internal standards.

Each set of arrows indicates the application of the equation shown to the left. Gray data points represent bins of 20 base pairs (bp). Labels above each graph are relate the variables in the equations to the values for which they are color-coded. Solid lines are moving averages over 25 bins (500 bp). (a) HydEn-seq end counts for reads mapping to the forward (black) and reverse (orange) strands from RNase H2-deficient pol1-L868M (α-LM; left), pol2-M644G (ɛ-MG; middle), and pol3-L612M (δ-LM; right) strains. (b) The observed strand bias, expressed as a log-ratio between forward- and reverse-strand end counts. (c) An approximation of the true replication strand bias log-ratio. Extra-replicative contributions were removed by comparing strains with opposite biases. The greater the biases in the chosen strains, the closer the approximation. (d) The fraction of replication events in which the bottom strand originates as the nascent lagging strand.

Supplementary Figure 3 Alkali-sensitive Okazaki fragment–sized DNA is not observed in the α-LM rnh201Δ or δ-LM rnh201D strains.

Purified genomic DNA was subjected to alkaline hydrolysis, alkaline-agarose electrophoresis and stained with SYBR® Gold Nucleic Acid Gel Stain, a sensitive fluorescent stain that can be used for detection of single strand DNA. The fraction of alkali-sensitive fragments was calculated by dividing the fluorescence intensity (arbitrary units) at each position along the gel by the total intensity for each lane. The experiment was performed in duplicate, and a representative gel image and quantitation is displayed. We note that Okazaki fragment-sized DNA (e.g. 150-300 base pair) fragments were not observed among the products of the alkaline hydrolysis of genomic DNA from the α-LM rnh201Δ or δ-LM rnh201Δ strains.

Supplementary Figure 4 Asymmetric phenotypes are associated with unrepaired ribonucleotides incorporated into nascent leading- versus lagging-strand DNA.

(a) Cell cycle progression is not affected by loss of RNH201 in the pol1-LM or the pol3-LM mutator strains with an increased number of unrepaired ribonucleotides in the nascent lagging strand. Cells were grown to mid-log phase at 30°C and processed for flow cytometry as described in13. The experiment was performed in duplicate and data are displayed as the mean % ± standard error. (b) The increase in total dNTP abundance conferred by loss of RNH201 is not significantly enhanced in the pol1-LM or the pol3-LM lagging strand polymerase mutator strains. Data for total dNTP abundance (dCTP, dATP, dTTP and dGTP) is displayed as the mean ± standard error. Each strain genotype was independently analyzed twice.

Supplementary Figure 5 URA3 mutation spectra for pol1 -L868M rnh201 Δ and pol3-L612M rnh201 Δ.

The coding strand of the 804 base pair URA3 open reading frame is shown. The sequence changes observed in independent ura3 mutants are depicted above the coding sequence for pol1-L868M rnh201∆ in red and below the coding sequence for pol3-L612M rnh201∆ in blue. Letters indicate single base substitutions, open triangles indicate single base deletions, and short lines above or below the coding sequence indicate multibase base deletions of between 2 and 5 base pairs. All strains had the URA3 reporter in OR2.

Supplementary information

Supplementary Figures and Text

Supplementary Figures 1–5 and Supplementary Tables 1–3 (PDF 5273 kb)

Supplementary Data Set 1

Original gel images for Figures 1–3 (PDF 28958 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Williams, J., Clausen, A., Lujan, S. et al. Evidence that processing of ribonucleotides in DNA by topoisomerase 1 is leading-strand specific. Nat Struct Mol Biol 22, 291–297 (2015). https://doi.org/10.1038/nsmb.2989

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nsmb.2989

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