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:

Ribose-seq: global mapping of ribonucleotides embedded in genomic DNA

An Addendum to this article was published on 19 July 2019

Abstract

Abundant ribonucleotide incorporation in DNA during replication and repair has profound consequences for genome stability, but the global distribution of ribonucleotide incorporation is unknown. We developed ribose-seq, a method for capturing unique products generated by alkaline cleavage of DNA at embedded ribonucleotides. High-throughput sequencing of these fragments in DNA from the yeast Saccharomyces cerevisiae revealed widespread ribonucleotide distribution, with a strong preference for cytidine and guanosine, and identified hotspots of ribonucleotide incorporation in nuclear and mitochondrial DNA. Ribonucleotides were primarily incorporated on the newly synthesized leading strand of nuclear DNA and were present upstream of (G+C)-rich tracts in the mitochondrial genome. Ribose-seq is a powerful tool for the systematic profiling of ribonucleotide incorporation in genomic DNA.

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: Ribose-seq method for mapping rNMPs in genomic DNA.
Figure 2: Distribution of rNMP incorporation in the S. cerevisiae genome.
Figure 3: Identity and sequence contexts of rNMP incorporation in S. cerevisiae genome.
Figure 4: Hotspots of rNMP incorporation in S. cerevisiae mitochondrial DNA, rDNA repeat and Ty1.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

References

  1. 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 

  2. Grossman, L.I., Watson, R. & Vinograd, J. The presence of ribonucleotides in mature closed-circular mitochondrial DNA. Proc. Natl. Acad. Sci. USA 70, 3339–3343 (1973).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. 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 

  4. Potenski, C.J. & Klein, H.L. How the misincorporation of ribonucleotides into genomic DNA can be both harmful and helpful to cells. Nucleic Acids Res. 42, 10226–10234 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Clausen, A.R., Zhang, S., Burgers, P.M., Lee, M.Y. & Kunkel, T.A. Ribonucleotide incorporation, proofreading and bypass by human DNA polymerase delta. DNA Repair (Amst.) 12, 121–127 (2013).

    Article  CAS  Google Scholar 

  6. Kasiviswanathan, R. & Copeland, W.C. Ribonucleotide discrimination and reverse transcription by the human mitochondrial DNA polymerase. J. Biol. Chem. 286, 31490–31500 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. 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 

  8. McDonald, J.P., Vaisman, A., Kuban, W., Goodman, M.F. & Woodgate, R. Mechanisms employed by Escherichia coli to prevent ribonucleotide incorporation into genomic DNA by pol V. PLoS Genet. 8, e1003030 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Zhu, H. & Shuman, S. Bacterial nonhomologous end joining ligases preferentially seal breaks with a 3′-OH monoribonucleotide. J. Biol. Chem. 283, 8331–8339 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Rumbaugh, J.A., Murante, R.S., Shi, S. & Bambara, R.A. Creation and removal of embedded ribonucleotides in chromosomal DNA during mammalian Okazaki fragment processing. J. Biol. Chem. 272, 22591–22599 (1997).

    Article  CAS  PubMed  Google Scholar 

  11. Randerath, K. et al. Formation of ribonucleotides in DNA modified by oxidative damage in vitro and in vivo. Characterization by 32P-postlabeling. Mutat. Res. 275, 355–366 (1992).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. 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 

  15. 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 

  16. 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 

  17. Yao, N.Y., Schroeder, J.W., Yurieva, O., Simmons, L.A. & O'Donnell, M.E. Cost of rNTP/dNTP pool imbalance at the replication fork. Proc. Natl. Acad. Sci. USA 110, 12942–12947 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Chiu, H.C. et al. RNA intrusions change DNA elastic properties and structure. Nanoscale 6, 10009–10017 (2014).

    Article  CAS  PubMed  Google Scholar 

  19. Caldecott, K.W. Molecular biology. Ribose–an internal threat to DNA. Science 343, 260–261 (2014).

    Article  CAS  PubMed  Google Scholar 

  20. 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 

  21. 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 

  22. 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 

  23. Crow, Y.J. et al. Mutations in genes encoding ribonuclease H2 subunits cause Aicardi-Goutieres syndrome and mimic congenital viral brain infection. Nat. Genet. 38, 910–916 (2006).

    Article  CAS  PubMed  Google Scholar 

  24. Schutz, K., Hesselberth, J.R. & Fields, S. Capture and sequence analysis of RNAs with terminal 2′,3′-cyclic phosphates. RNA 16, 621–631 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Remus, B.S. & Shuman, S. Distinctive kinetics and substrate specificities of plant and fungal tRNA ligases. RNA 20, 462–473 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Cooper, D.A., Jha, B.K., Silverman, R.H., Hesselberth, J.R. & Barton, D.J. Ribonuclease L and metal-ion-independent endoribonuclease cleavage sites in host and viral RNAs. Nucleic Acids Res. 42, 5202–5216 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Krokan, H.E., Drablos, F. & Slupphaug, G. Uracil in DNA–occurrence, consequences and repair. Oncogene 21, 8935–8948 (2002).

    Article  CAS  PubMed  Google Scholar 

  28. Lindahl, T., Ljungquist, S., Siegert, W., Nyberg, B. & Sperens, B. DNA N-glycosidases: properties of uracil-DNA glycosidase from Escherichia coli. J. Biol. Chem. 252, 3286–3294 (1977).

    Article  CAS  PubMed  Google Scholar 

  29. Fasullo, M., Tsaponina, O., Sun, M. & Chabes, A. Elevated dNTP levels suppress hyper-recombination in Saccharomyces cerevisiae S-phase checkpoint mutants. Nucleic Acids Res. 38, 1195–1203 (2010).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed Central  Google Scholar 

  31. Foury, F., Roganti, T., Lecrenier, N. & Purnelle, B. The complete sequence of the mitochondrial genome of Saccharomyces cerevisiae. FEBS Lett. 440, 325–331 (1998).

    Article  CAS  PubMed  Google Scholar 

  32. Gerhold, J.M., Aun, A., Sedman, T., Joers, P. & Sedman, J. Strand invasion structures in the inverted repeat of Candida albicans mitochondrial DNA reveal a role for homologous recombination in replication. Mol. Cell 39, 851–861 (2010).

    Article  CAS  PubMed  Google Scholar 

  33. Yabuki, N., Terashima, H. & Kitada, K. Mapping of early firing origins on a replication profile of budding yeast. Genes Cells 7, 781–789 (2002).

    Article  CAS  PubMed  Google Scholar 

  34. Moraes, C.T. What regulates mitochondrial DNA copy number in animal cells? Trends Genet. 17, 199–205 (2001).

    Article  CAS  PubMed  Google Scholar 

  35. Szostak, J.W. & Wu, R. Unequal crossing over in the ribosomal DNA of Saccharomyces cerevisiae. Nature 284, 426–430 (1980).

    Article  CAS  PubMed  Google Scholar 

  36. Hani, J. & Feldmann, H. tRNA genes and retroelements in the yeast genome. Nucleic Acids Res. 26, 689–696 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Mieczkowski, P.A., Lemoine, F.J. & Petes, T.D. Recombination between retrotransposons as a source of chromosome rearrangements in the yeast Saccharomyces cerevisiae. DNA Repair (Amst.) 5, 1010–1020 (2006).

    Article  CAS  Google Scholar 

  38. El Hage, A., Webb, S., Kerr, A. & Tollervey, D. Genome-wide distribution of RNA-DNA hybrids identifies RNase H targets in tRNA genes, retrotransposons and mitochondria. PLoS Genet. 10, e1004716 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  39. Barrett, T. et al. NCBI GEO: archive for functional genomics data sets–update. Nucleic Acids Res. 41, D991–D995 (2013).

    Article  CAS  PubMed  Google Scholar 

  40. Clausen, A.R. et al. Tracking replication enzymology in vivo by genome-wide mapping of ribonucleotide incorporation. Nat. Struct. Molec. Biol. doi:10.1038/nsmb.2957 (26 January 2015).

  41. Reijns, M.A.M. et al. Lagging strand replication shapes the mutational landscape of the genome. Nature doi:10.1038/nature14183 (26 January 2015).

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

  43. Clark, A.B. et al. Functional analysis of human MutSα and MutSβ complexes in yeast. Nucleic Acids Res. 27, 736–742 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Morrison, A., Bell, J.B., Kunkel, T.A. & Sugino, A. Eukaryotic DNA polymerase amino acid sequence required for 3′ → 5′ exonuclease activity. Proc. Natl. Acad. Sci. USA 88, 9473–9477 (1991).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Jin, Y.H. et al. The 3′→5′ exonuclease of DNA polymerase δ can substitute for the 5′ flap endonuclease Rad27/Fen1 in processing Okazaki fragments and preventing genome instability. Proc. Natl. Acad. Sci. USA 98, 5122–5127 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Storici, F., Bebenek, K., Kunkel, T.A., Gordenin, D.A. & Resnick, M.A. RNA-templated DNA repair. Nature 447, 338–341 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Kokoska, R.J., McCulloch, d. & Kunkel, T.A. The efficiency and specificity of apurinic/apyrimidinic site bypass by human DNA polymerase η and Sulfolobus solfataricus Dpo4. J. Biol. Chem. 278, 50537–50545 (2003).

    Article  CAS  PubMed  Google Scholar 

  49. Kuchta, R.D. & Stengel, G. Mechanism and evolution of DNA primases. Biochim. Biophys. Acta 1804, 1180–1189 (2010).

    Article  CAS  PubMed  Google Scholar 

  50. Smith, C.W.J. RNA-Protein Interactions: A Practical Approach 1st edn. (Oxford Univ. Press, New York, 1998).

  51. Lhomme, J., Constant, J.F. & Demeunynck, M. Abasic DNA structure, reactivity, and recognition. Biopolymers 52, 65–83 (1999).

    Article  CAS  PubMed  Google Scholar 

  52. Bailly, V. & Verly, W.G. Possible roles of β-elimination and δ-elimination reactions in the repair of DNA containing AP (apurinic/apyrimidinic) sites in mammalian cells. Biochem. J. 253, 553–559 (1988).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Quinlan, A.R. & Hall, I.M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Dale, R.K., Pedersen, B.S. & Quinlan, A.R. Pybedtools: a flexible Python library for manipulating genomic data sets and annotations. Bioinformatics 27, 3423–3424 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Siow, C.C., Nieduszynska, S.R., Muller, C.A. & Nieduszynski, C.A. OriDB, the DNA replication origin database updated and extended. Nucleic Acids Res. 40, D682–D686 (2012).

    Article  CAS  PubMed  Google Scholar 

  56. Zhang, Y. et al. Model-based analysis of ChIP-Seq (MACS). Genome Biol. 9, R137 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  57. Sokal, R.R. & Rohlf, F.J. Biometry: The Principles and Practice of Statistics in Biological Research 3rd edn. (W.H. Freeman, New York, 1995).

Download references

Acknowledgements

We thank N.V. Hud and L.D. Williams for support with urea-PAGE gels and for advice on this study and the manuscript; M. Goodman, B. Weiss and I.K. Jordan for suggestions on this study and the manuscript and assistance with data analysis; S. Garrey for AtRNL and Tpt1 protein purification; C. Cox for sequencing; Y. Shen for assistance with statistical analysis; A. Gombolay and L. Shetty for technical help; and all members of the Storici laboratory for advice in the course of the study. This research was supported by US National Science Foundation award number MCB-1021763 (to F.S.), Georgia Research Alliance award number R9028 (to F.S.), an American Cancer Society Research Scholar Grant (to J.R.H.), a Damon Runyon-Rachleff Innovation Award from the Damon Runyon Cancer Research Foundation (to J.R.H.) and the University of Colorado Golfers Against Cancer (to J.R.H.).

Author information

Authors and Affiliations

Authors

Contributions

K.D.K. conducted all of the experiments for in vitro biochemical assays and ribose-seq library construction, and most of the experiments for yeast in vivo DSB repair assay with oligonucleotides, with assistance from S.B. J.R.H. conducted the sequencing analysis. F.S., together with K.D.K. and J.R.H., designed experiments, assisted in data analysis and wrote the manuscript.

Corresponding authors

Correspondence to Jay R Hesselberth or Francesca Storici.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Mechanism of alkaline cleavage of ribonucleotides in DNA

The ribonucleoside embedded in double-stranded DNA is in red. During alkaline treatment, DNA strands are denatured, and cleavage occurs at the rNMP site, generating a 2′,3′-cyclic phosphate end and an opposite 5′-hydroxyl end. The 2′,3′-cyclic phosphate is in equilibrium with 2′-phosphate and 3′-phosphate forms. Boxes in black indicate the 2′,3′-cyclic phosphate and 2′-phosphate DNA termini, which are substrates of AtRNL.

Supplementary Figure 2 3′ base bias for AtRNL ligation

Hot 5′-radiolabeled 30-nt DNA oligo with a single rNMP (either A, G, U, or C) in the 22nd position was mixed with cold equimolar 30-nt DNA oligos with rNMPs of 3 other bases in the 22nd positions. 5′-radiolabel is indicated by ‘P’ in purple. The mixture was treated with 0.3M NaOH for 2 hr at 55 °C and neutralized. 100 nM of alkali-cleaved products (25 nM of each base) were then incubated with 1 µM AtRNL in appropriate buffer (see Methods) for 1 hr at 30 °C. The resulting products were treated with T5 exonuclease for 2 hr at 37 °C. Aliquots were withdrawn after appropriate steps and quenched. The products were analyzed by urea-PAGE. The circular 22-mer migrates faster than the unligated, linear 22-mer. Only circular products were resistant to T5 exonuclease while all linear substrates/products were degraded. Median percentages of circular 22-mer formation from four independent reactions are displayed. See Supplementary Table 1 for more statistics. First left lane, ss DNA ladder. No 3′ base bias was observed for AtRNL ligation (see Supplementary Table 1). Self-ligation was preferred to dimerization with a shorter 22-nt substrate; however, with the shorter substrate, lower levels of linear dimers, which are not resistant to T5 exonuclease, and circular dimers were observed. Increasing the length of the ss DNA substrate from 22 nt to 32 nt eliminated dimerization (Fig. 1a).

Supplementary Figure 3 Ribose-seq library from genomic DNA of S. cerevisiae rnh201Δ (KK-100) cells

Appropriate PCR products were analyzed by PAGE. ‘P’ indicates primers-only. No amplification product was observed when either (a) AtRNL ligation step or (b) alkali treatment was omitted. Tpt1 denotes the step of 2’-phosphate removal at the ligation junction in Figure 1a. First left lane, ds DNA ladder.

Supplementary Figure 4 Bypass of a single rNMP by Phusion DNA polymerase

5′-radiolabeled 30-nt primer, ByPrim (Supplementary Table 8), was annealed to the 46-nt template oligo containing either rCMP (ByTemp.rC) or rUMP (ByTemp.rU) in the 8th position. 100 nM of annealed substrate was incubated with 0.2 units of Phusion High-Fidelity DNA Polymerase (NEB) and 2 mM dNTPs in appropriate buffer (see Methods) for 30 sec at 72 °C. The reactions were quenched and analyzed by urea-PAGE. Median bypass probabilities from four independent reactions are shown. See Supplementary Table 5 for more statistics. First left lane, ss DNA ladder. The primer extension assay showed no significant difference between bypass efficiency over rUMP and rCMP by Phusion DNA polymerase (Supplementary Table 5).

Supplementary Figure 5 Normalized frequency of nucleotides surrounding the rNMP sites

Normalized frequency of nucleotides relative to (a) nuclear and (b) mitochondrial mapped positions of sequences from ribose-seq library, PCR-amplified with EconoTaq DNA Polymerase (Lucigen), of genomic DNA from S. cerevisiae rnh201Δ (KK-100) cells. Position 0 corresponds to the rNMP. Negative and positive numbers (from -10 to -1 and 1 to 10) correspond to upstream and downstream positions from the rNMP, respectively. Frequencies were normalized to either nuclear or mitochondrial genomic mononucleotide frequencies. Normalized frequency of nucleotides relative to (c) nuclear and (d) mitochondrial mapped positions of sequences from ribose-seq library of genomic DNA from S. cerevisiae rnh201Δ (KK-30) cells. Normalized frequency of nucleotides relative to (e) nuclear and (f) mitochondrial mapped positions of sequences from ribose-seq library of genomic DNA from S. cerevisiae rnh1Δ rnh201Δ (KK-174) cells. Normalized frequency of nucleotides relative to (g) nuclear and (h) mitochondrial mapped positions of sequences from ribose-seq library of genomic DNA from S. cerevisiae rnh1Δ rnh201Δ (KK-125) cells. Normalized frequency of nucleotides relative to (i) nuclear and (j) mitochondrial mapped positions of sequences from ribose-seq library of genomic DNA from S. cerevisiae rnh1Δ rnh201Δ ung1Δ (KK-164) cells. Normalized frequency of nucleotides relative to (k) nuclear and (l) mitochondrial mapped positions of sequences from ribose-seq library of genomic DNA from S. cerevisiae pol2-M644G rnh201Δ (KK-170) cells. Normalized frequency of nucleotides relative to (m) nuclear and (n) mitochondrial mapped positions of sequences from ribose-seq library of genomic DNA from S. cerevisiae pol3-5DV rnh201Δ (KK-120) cells.

Supplementary Figure 6 Zoom-out of normalized frequency of nucleotides surrounding the rNMP sites

Normalized frequency of nucleotides relative to (a) nuclear and (b) mitochondrial mapped positions of sequences from ribose-seq library, PCR-amplified with EconoTaq DNA Polymerase (Lucigen), of genomic DNA from S. cerevisiae rnh201Δ (KK-100) cells. Position 0 corresponds to the rNMP. Negative and positive numbers (from -100 to -1 and 1 to 100) correspond to upstream and downstream positions from the rNMP, respectively. Frequencies were normalized to either nuclear or mitochondrial genomic mononucleotide frequencies. Normalized frequency of nucleotides relative to (c) nuclear and (d) mitochondrial mapped positions of sequences from ribose-seq library of genomic DNA from S. cerevisiae rnh201Δ (KK-30) cells. Normalized frequency of nucleotides relative to (e) nuclear and (f) mitochondrial mapped positions of sequences from ribose-seq library of genomic DNA from S. cerevisiae rnh1Δ rnh201Δ (KK-174) cells. Normalized frequency of nucleotides relative to (g) nuclear and (h) mitochondrial mapped positions of sequences from ribose-seq library of genomic DNA from S. cerevisiae rnh1Δ rnh201Δ (KK-125) cells. Normalized frequency of nucleotides relative to (i) nuclear and (j) mitochondrial mapped positions of sequences from ribose-seq library of genomic DNA from S. cerevisiae rnh1Δ rnh201Δ ung1Δ (KK-164) cells. Normalized frequency of nucleotides relative to (k) nuclear and (l) mitochondrial mapped positions of sequences from ribose-seq library of genomic DNA from S. cerevisiae pol2-M644G rnh201Δ (KK-170) cells. Normalized frequency of nucleotides relative to (m) nuclear and (n) mitochondrial mapped positions of sequences from ribose-seq library of genomic DNA from S. cerevisiae pol2-4 rnh201Δ (KK-107) cells. Normalized frequency of nucleotides relative to (o) nuclear and (p) mitochondrial mapped positions of sequences from ribose-seq library of genomic DNA from S. cerevisiae pol3-5DV rnh201Δ (KK-120) cells.

Supplementary Figure 7 Targeting of rGMP and rUMP by RNase H2 and uracil DNA N-glycosylase during DSB repair in S. cerevisiae cells

a, Diagram and sequence of the chromosomal leu2 region targeted by DNA-control LEU2.D, rGMP-containing LEU2.rG, dUMP-containing LEU2.dU, and rUMP-containing LEU2.rU oligos (Supplementary Table 8). StuI recognition sequence is underlined in turquoise. Position of either rGMP, dUMP, or rUMP was selected so that it is about 4-5 nt upstream of the G-T mispair. Both RNase H2-initiated excision repair (RER) and base excision repair (BER) remove a short ss DNA region downstream of the damage during the repair13,27. b, The oligos were transformed to either RNase H2- and uracil DNA N-glycosylase-proficient wild-type (WT; FRO-767,768), RNase H2-deficient (rnh201; FRO-984,985), or DNA N-glycosylase-deficient (ung1; KK-158,159) S. cerevisiae cells (see Supplementary Table 2). Median percentages of StuI-cut Leu+ transformants from four independent transformations are shown with ranges as bars. For each transformation, 20 Leu+ transformants were selected for analysis. Mann-Whitney U-test was implemented for statistical analysis against the WT. P values of less than 0.05 are marked as asterisk. See Supplementary Table 6 for more statistics.

Supplementary Figure 8 Normalized frequency of nucleotides surrounding the rNMP sites on leading and lagging strands

Normalized frequency of nucleotides relative to mapped positions of sequences in (a) leading and (b) lagging strands from ribose-seq library of genomic DNA from S. cerevisiae rnh1Δ rnh201Δ (KK-174) cells. Position 0 corresponds to the rNMP. Negative and positive numbers (from -10 to -1 and 1 to 10) correspond to upstream and downstream positions from the rNMP, respectively. ARSs with Trep of no longer than 25 min were selected with flanking size of 10 kb. Frequencies were normalized to genomic mononucleotide frequencies of either leading or lagging strand of the selected ARSs and flanking size. Normalized frequency of nucleotides relative to mapped positions of sequences in (c) leading and (d) lagging strands from ribose-seq library of genomic DNA from S. cerevisiae pol3-5DV rnh201Δ (KK-120) cells.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–8 and Supplementary Tables 1–8 (PDF 4491 kb)

Supplementary Data

Results of peak calling analysis of all ribose-seq libraries in this study. (XLS 135 kb)

Supplementary Software

Modmap software version 1.1 includes the pipeline for sequencing analysis (using bowtie, samtools and bedtools) and a series of bash, Python and R scripts for analyzing data and generating plots. The software is available here as a ZIP file and is maintained at http://github.com/hesselberthlab/modmap/. (ZIP 360 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Koh, K., Balachander, S., Hesselberth, J. et al. Ribose-seq: global mapping of ribonucleotides embedded in genomic DNA. Nat Methods 12, 251–257 (2015). https://doi.org/10.1038/nmeth.3259

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nmeth.3259

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research