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:

RTEL1 suppresses G-quadruplex-associated R-loops at difficult-to-replicate loci in the human genome

Abstract

Oncogene activation during tumorigenesis generates DNA replication stress, a known driver of genome rearrangements. In response to replication stress, certain loci, such as common fragile sites and telomeres, remain under-replicated during interphase and subsequently complete locus duplication in mitosis in a process known as ‘MiDAS’. Here, we demonstrate that RTEL1 (regulator of telomere elongation helicase 1) has a genome-wide role in MiDAS at loci prone to form G-quadruplex-associated R-loops, in a process that is dependent on its helicase function. We reveal that SLX4 is required for the timely recruitment of RTEL1 to the affected loci, which in turn facilitates recruitment of other proteins required for MiDAS, including RAD52 and POLD3. Our findings demonstrate that RTEL1 is required for MiDAS and suggest that RTEL1 maintains genome stability by resolving conflicts that can arise between the replication and transcription machineries.

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: RTEL1 is essential for MiDAS following replication stress.
Fig. 2: RTEL1 is essential for the maintenance of genome stability.
Fig. 3: RTEL1 prevents R-loop accumulation in interphase cells.
Fig. 4: DRIP-seq analysis reveals that sequences enriched in RTEL1-depleted or APH-treated cells share similar features.
Fig. 5: RTEL1 depletion induces an accumulation of G4s, which inhibits cell growth.
Fig. 6: MEF cells expressing ATPase-dead (K48R) Rtel1 have increased R-loops in interphase and reduced MiDAS in prometaphase.
Fig. 7: RTEL1 acts downstream of SLX4 in the MiDAS pathway.
Fig. 8: Overexpression of RNase H1 can remove R-loops, and partially blocks MiDAS under RS conditions.

Similar content being viewed by others

Data availability

The raw sequence data for each DRIP-seq sample are deposited at NCBI Bioproject (https://www.ncbi.nlm.nih.gov/bioproject/PRJNA510030/). The names of the samples are YL_sample_7.fq.gz (siRNA-control), YL_sample_8.fq.gz (siRNA-control+RNase H), YL_sample_9.fq.gz (siRNA-RTEL1), YL_sample_10.fq.gz (siRNA-RTEL1+RNase H), YL_sample_11.fq.gz (siRNA-control+APH) and YL_sample_12.fq.gz (siRNA-control+APH+RNase H). The final processed data are deposited at the NCBI GEO database (accession no. GSE129907). The names of the samples are: sample_7.bed (siRNA-control), sample_8.bed (siRNA-control+RNase H), sample_9.bed (siRNA-RTEL1), sample_10.bed (siRNA-RTEL1+RNase H), sample_11.bed (siRNA-control+APH) and sample_12.bed (siRNA-control+APH+RNase H). The source data for graphs, western blots and DNA gels are available with the paper online. The source data for all of the images in figures are deposited at Figshare (https://doi.org/10.6084/m9.figshare.11833368).

Code availability

Custom-made Bash and Python scripts are available upon request from the corresponding authors.

References

  1. Baddeley, D. et al. Measurement of replication structures at the nanometer scale using super-resolution light microscopy. Nucleic Acids Res 38, e8 (2010).

    CAS  PubMed  Google Scholar 

  2. Arlt, M. F., Durkin, S. G., Ragland, R. L. & Glover, T. W. Common fragile sites as targets for chromosome rearrangements. DNA Repair 5, 1126–1135 (2006).

    CAS  PubMed  Google Scholar 

  3. Burrell, R. A. et al. Replication stress links structural and numerical cancer chromosomal instability. Nature 494, 492–496 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Burrow, A. A., Williams, L. E., Pierce, L. C. & Wang, Y. H. Over half of breakpoints in gene pairs involved in cancer-specific recurrent translocations are mapped to human chromosomal fragile sites. BMC Genomics 10, 59 (2009).

    PubMed  PubMed Central  Google Scholar 

  5. Halazonetis, T. D., Gorgoulis, V. G. & Bartek, J. An oncogene-induced DNA damage model for cancer development. Science 319, 1352–1355 (2008).

    CAS  PubMed  Google Scholar 

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

    PubMed  Google Scholar 

  7. Dilley, R. L. et al. Break-induced telomere synthesis underlies alternative telomere maintenance. Nature 539, 54–58 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Min, J., Wright, W. E. & Shay, J. W. Alternative lengthening of telomeres mediated by mitotic DNA synthesis engages break-induced replication processes. Mol. Cell. Biol. 37, e00226-17 (2017).

    PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  10. Özer, Ö., Bhowmick, R., Liu, Y. & Hickson, I. D. Human cancer cells utilize mitotic DNA synthesis to resist replication stress at telomeres regardless of their telomere maintenance mechanism. Oncotarget 9, 15836–15846 (2018).

    PubMed  PubMed Central  Google Scholar 

  11. Bhowmick, R., Minocherhomji, S. & Hickson, I. D. RAD52 facilitates mitotic DNA synthesis following replication stress. Mol. Cell 64, 1117–1126 (2016).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  14. 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).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  16. Stirling, P. C. & Hieter, P. Canonical DNA repair pathways influence R-loop-driven genome instability. J. Mol. Biol. 429, 3132–3138 (2017).

    CAS  PubMed  Google Scholar 

  17. Ginno, P. A., Lott, P. L., Christensen, H. C., Korf, I. & Chedin, F. R-loop formation is a distinctive characteristic of unmethylated human CpG Island promoters. Mol. Cell 45, 814–825 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Fungtammasan, A., Walsh, E., Chiaromonte, F., Eckert, K. A. & Makova, K. D. A genome-wide analysis of common fragile sites: what features determine chromosomal instability in the human genome? Genome Res. 22, 993–1005 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Kuznetsov, V. A., Bondarenko, V., Wongsurawat, T., Yenamandra, S. P. & Jenjaroenpun, P. Toward predictive R-loop computational biology: genome-scale prediction of R-loops reveals their association with complex promoter structures, G-quadruplexes and transcriptionally active enhancers. Nucleic Acids Res. 46, 7566–7585 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Loomis, E. W., Sanz, L. A., Chedin, F. & Hagerman, P. J. Transcription-associated R-loop formation across the human FMR1 CGG-repeat region. PLoS Genet. 10, e1004294 (2014).

    PubMed  PubMed Central  Google Scholar 

  21. Pedersen, R. T., Kruse, T., Nilsson, J., Oestergaard, V. H. & Lisby, M. TopBP1 is required at mitosis to reduce transmission of DNA damage to G1 daughter cells. J. Cell Biol. 210, 565–582 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Vannier, J. B., Pavicic-Kaltenbrunner, V., Petalcorin, M. I., Ding, H. & Boulton, S. J. RTEL1 dismantles T loops and counteracts telomeric G4-DNA to maintain telomere integrity. Cell 149, 795–806 (2012).

    CAS  PubMed  Google Scholar 

  23. Sotiriou, S. K. et al. Mammalian RAD52 functions in break-induced replication repair of collapsed DNA replication forks. Mol. Cell 64, 1127–1134 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Chan, Y. W., Fugger, K. & West, S. C. Unresolved recombination intermediates lead to ultra-fine anaphase bridges, chromosome breaks and aberrations. Nat. Cell Biol. 20, 92–103 (2018).

    CAS  PubMed  Google Scholar 

  25. Liu, Y., Nielsen, C. F., Yao, Q. & Hickson, I. D. The origins and processing of ultra fine anaphase DNA bridges. Curr. Opin. Genet. Dev. 26, 1–5 (2014).

    PubMed  Google Scholar 

  26. Nielsen, C. F. et al. PICH promotes sister chromatid disjunction and co-operates with topoisomerase II in mitosis. Nat. Commun. 6, 8962 (2015).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  28. Arora, R. et al. RNaseH1 regulates TERRA-telomeric DNA hybrids and telomere maintenance in ALT tumour cells. Nat. Commun. 5, 5220 (2014).

    CAS  PubMed  Google Scholar 

  29. Hall, A. C., Ostrowski, L. A., Pietrobon, V. & Mekhail, K. Repetitive DNA loci and their modulation by the non-canonical nucleic acid structures R-loops and G-quadruplexes. Nucleus 8, 162–181 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Barber, L. J. et al. RTEL1 maintains genomic stability by suppressing homologous recombination. Cell 135, 261–271 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Boguslawski, S. J. et al. Characterization of monoclonal-antibody to DNA · RNA and its application to immunodetection of hybrids. J. Immunol. Methods 89, 123–130 (1986).

    CAS  PubMed  Google Scholar 

  33. Chan, K. L., Palmai-Pallag, T., Ying, S. M. & Hickson, I. D. Replication stress induces sister-chromatid bridging at fragile site loci in mitosis. Nat. Cell Biol. 11, 753–760 (2009).

    CAS  PubMed  Google Scholar 

  34. Fan, Q., Zhang, F., Barrett, B., Ren, K. Q. & Andreassen, P. R. A role for monoubiquitinated FANCD2 at telomeres in ALT cells. Nucleic Acids Res. 37, 1740–1754 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Howlett, N. G., Taniguchi, T., Durkin, S. G., D’Andrea, A. D. & Glover, T. W. The Fanconi anemia pathway is required for the DNA replication stress response and for the regulation of common fragile site stability. Hum. Mol. Genet. 14, 693–701 (2005).

    CAS  PubMed  Google Scholar 

  36. Sordet, O. et al. Ataxia telangiectasia mutated activation by transcription- and topoisomerase I-induced DNA double-strand breaks. EMBO Rep. 10, 887–893 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Tresini, M. et al. The core spliceosome as target and effector of non-canonical ATM signalling. Nature 523, 53–58 (2015).

    PubMed Central  Google Scholar 

  38. Okamoto, Y. et al. Replication stress induces accumulation of FANCD2 at central region of large fragile genes. Nucleic Acids Res. 46, 2932–2944 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Glover, T. W., Berger, C., Coyle, J. & Echo, B. DNA polymerase-α inhibition by aphidicolin induces gaps and breaks at common fragile sites in human chromosomes. Hum. Genet. 67, 136–142 (1984).

    CAS  PubMed  Google Scholar 

  40. Kumar, R. et al. HumCFS: a database of fragile sites in human chromosomes. BMC Genomics 19, 985 (2019).

    PubMed  PubMed Central  Google Scholar 

  41. Mrasek, K. et al. Global screening and extended nomenclature for 230 aphidicolin-inducible fragile sites, including 61 yet unreported ones. Int. J. Oncol. 36, 929–940 (2010).

    PubMed  Google Scholar 

  42. Forbes, S. A. et al. COSMIC: somatic cancer genetics at high-resolution. Nucleic Acids Res. 45, D777–D783 (2017).

    CAS  PubMed  Google Scholar 

  43. Kikin, O., D’Antonio, L. & Bagga, P. S. QGRS Mapper: a web-based server for predicting G-quadruplexes in nucleotide sequences. Nucleic Acids Res. 34, W676–W682 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Hansel-Hertsch, R., Spiegel, J., Marsico, G., Tannahill, D. & Balasubramanian, S. Genome-wide mapping of endogenous G-quadruplex DNA structures by chromatin immunoprecipitation and high-throughput sequencing. Nat. Protoc. 13, 551–564 (2018).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Gavathiotis, E., Heald, R. A., Stevens, M. F. & Searle, M. S. Drug recognition and stabilisation of the parallel-stranded DNA quadruplex d(TTAGGGT)4 containing the human telomeric repeat. J. Mol. Biol. 334, 25–36 (2003).

    CAS  PubMed  Google Scholar 

  47. Gowan, S. M., Heald, R., Stevens, M. F. & Kelland, L. R. Potent inhibition of telomerase by small-molecule pentacyclic acridines capable of interacting with G-quadruplexes. Mol. Pharmacol. 60, 981–988 (2001).

    CAS  PubMed  Google Scholar 

  48. Heald, R. A. et al. Antitumor polycyclic acridines. 8.1 Synthesis and telomerase-inhibitory activity of methylated pentacyclic acridinium salts. J. Med. Chem. 45, 590–597 (2002).

    CAS  PubMed  Google Scholar 

  49. Wu, X., Sandhu, S., Nabi, Z. & Ding, H. Generation of a mouse model for studying the role of upregulated RTEL1 activity in tumorigenesis. Transgenic Res. 21, 1109–1115 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Yamamoto, K. N. et al. Involvement of SLX4 in interstrand cross-link repair is regulated by the Fanconi anemia pathway. Proc. Natl Acad. Sci. USA 108, 6492–6496 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Wilson, J. S. et al. Localization-dependent and -independent roles of SLX4 in regulating telomeres. Cell Rep. 4, 853–860 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Parajuli, S. et al. Human ribonuclease H1 resolves R-loops and thereby enables progression of the DNA replication fork. J. Biol. Chem. 292, 15216–15224 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Skourti-Stathaki, K., Proudfoot, N. J. & Gromak, N. Human senataxin resolves RNA/DNA hybrids formed at transcriptional pause sites to promote Xrn2-dependent termination. Mol. Cell 42, 794–805 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Yuce, O. & West, S. C. Senataxin, defective in the neurodegenerative disorder ataxia with oculomotor apraxia 2, lies at the interface of transcription and the DNA damage response. Mol. Cell. Biol. 33, 406–417 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Zhao, H. C., Zhu, M., Limbo, O. & Russell, P. RNase H eliminates R-loops that disrupt DNA replication but is nonessential for efficient DSB repair. EMBO Rep. 19, e45335 (2018).

    PubMed  PubMed Central  Google Scholar 

  56. Cohen, S. et al. Senataxin resolves RNA:DNA hybrids forming at DNA double-strand breaks to prevent translocations. Nat. Commun. 9, 533 (2018).

    PubMed  PubMed Central  Google Scholar 

  57. Aguilera, A. & Garcia-Muse, T. R Loops: from transcription byproducts to threats to genome stability. Mol. Cell 46, 115–124 (2012).

    CAS  PubMed  Google Scholar 

  58. Hamperl, S. & Cimprich, K. A. The contribution of co-transcriptional RNA:DNA hybrid structures to DNA damage and genome instability. DNA Repair 19, 84–94 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Harrigan, J. A. et al. Replication stress induces 53BP1-containing OPT domains in G1 cells. J. Cell Biol. 193, 97–108 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Lemmens, B. et al. DNA replication determines timing of mitosis by restricting CDK1 and PLK1 activation. Mol. Cell 71, 117–128 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Munoz, I. M. et al. Coordination of structure-specific nucleases by human SLX4/BTBD12 is required for DNA repair. Mol. Cell 35, 116–127 (2009).

    CAS  PubMed  Google Scholar 

  63. Takedachi, A. SLX4 interacts with RTEL1 to prevent transcription-mediated DNA replication perturbations Nat. Struct. Mol. Biol. https://doi.org/10.1038/s41594-020-0419-3 (2020).

  64. Ballew, B. J. et al. Germline mutations of regulator of telomere elongation helicase 1, RTEL1, in dyskeratosis congenita. Hum. Genet. 132, 473–480 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Ballew, B. J. et al. A recessive founder mutation in regulator of telomere elongation helicase 1, RTEL1, underlies severe immunodeficiency and features of Hoyeraal Hreidarsson syndrome. PLoS Genet. 9, e1003695 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Deng, Z. et al. Inherited mutations in the helicase RTEL1 cause telomere dysfunction and Hoyeraal–Hreidarsson syndrome. Proc. Natl Acad. Sci. USA 110, E3408–E3416 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Egan, K. M. et al. Cancer susceptibility variants and the risk of adult glioma in a US case–control study. J. Neurooncol. 104, 535–542 (2011).

    PubMed  PubMed Central  Google Scholar 

  68. Liu, Y. et al. Polymorphisms of LIG4, BTBD2, HMGA2 and RTEL1 genes involved in the double-strand break repair pathway predict glioblastoma survival. J. Clin. Oncol. 28, 2467–2474 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Rajaraman, P. et al. Genome-wide association study of glioma and meta-analysis. Hum. Genet. 131, 1877–1888 (2012).

    PubMed  PubMed Central  Google Scholar 

  70. Shete, S. et al. Genome-wide association study identifies five susceptibility loci for glioma. Nat. Genet. 41, 899–904 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Wrensch, M. et al. Variants in the CDKN2B and RTEL1 regions are associated with high-grade glioma susceptibility. Nat. Genet. 41, 905–908 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Bai, C. et al. Overexpression of M68/DcR3 in human gastrointestinal tract tumors independent of gene amplification and its location in a four-gene cluster. Proc. Natl Acad. Sci. USA 97, 1230–1235 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Muleris, M., Almeida, A., Gerbaultseureau, M., Malfoy, B. & Dutrillaux, B. Identification of amplified DNA sequences in breast cancer and their organization within homogeneously staining regions. Genes Chromosomes Cancer 14, 155–163 (1995).

    CAS  PubMed  Google Scholar 

  74. Pitti, R. M. et al. Genomic amplification of a decoy receptor for Fas ligand in lung and colon cancer. Nature 396, 699–703 (1998).

    CAS  PubMed  Google Scholar 

  75. Garribba, L. et al. Inducing and detecting mitotic DNA synthesis at difficult-to-replicate oci. Methods Enzymol. 601, 45–58 (2018).

    CAS  PubMed  Google Scholar 

  76. Bjerregaard, V. A., Garribba, L., McMurray, C. T., Hickson, I. D. & Liu, Y. Folate deficiency drives mitotic missegregation of the human FRAXA locus. Proc. Natl Acad. Sci. USA 115, 13003–13008 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Bizard, A. H., Nielsen, C. F. & Hickson, I. D. Detection of ultrafine anaphase bridges. Methods Mol. Biol. 1672, 495–508 (2018).

    CAS  PubMed  Google Scholar 

  78. Skourti-Stathaki, K., Kamieniarz-Gdula, K. & Proudfoot, N. J. R-loops induce repressive chromatin marks over mammalian gene terminators. Nature 516, 436–439 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  80. Schwab, R. A. et al. The Fanconi anemia pathway maintains genome stability by coordinating replication and transcription. Mol. Cell 60, 351–361 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank members of the Hickson and Liu groups for helpful discussions, and J.P. Diaz and E. Hoffmann for critical reading of the manuscript. We also thank W. Qiao and V. Lee (BGI Hong Kong) for processing the DRIP-seq samples. This work was supported by grants from the Chinese Scholarship Council (PhD fellowship 201406170048; W.W.), the Danish Medical Research Council (Postdoc fellowship DFF-4004-00155B; R.B.), the European Commission (FP7 Marie Curie Fellowship; Ö.Ö.), US NIH grants (GM56888 and U54 OD020355; J.H.P.), US NCI MSK Cancer Center core grant P30 CA008748 (J.H.P.), The Nordea Foundation of Denmark (I.D.H.), The Danish National Research Foundation (DNRF115; I.D.H. and Y.L.) and the European Commission (H2020/ Marie Skłodowska-Curie, 859853; I.D.H. and Y.L.).

Author information

Authors and Affiliations

Authors

Contributions

W.W., R.B., Ö.Ö., F.G., R.S.T., E.S., P.H.R. and L.R. carried out experiments. I.V. performed analysis of bioinformatic data. W.W., R.B., J.H.P., I.D.H. and Y.L. designed experiments and interpreted results. I.D.H. and Y.L. wrote the manuscript and all authors edited it.

Corresponding authors

Correspondence to Ian D. Hickson or Ying Liu.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Beth Moorefield was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

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

Extended data

Extended Data Fig. 1 RTEL1 depletion by 3’UTR siRNAs causes R-loop accumulation in interphase and reduced MiDAS in prometaphase cells under replication stress conditions.

a, Experimental workflow for the analysis of MiDAS in prometaphase cells in U2OS cells following RTEL1 depletion and APH treatment. b & c, Representative images of Immunoblotting analysis (b) and quantification (c) of RTEL1 expression following transfection with a control and two 3’UTR targeting siRNA. GAPDH was used as a loading control. The quantification of relative protein levels was performed using Fiji/ImageJ software. Error bars indicate mean ± s.d. (N = 3). **P < 0.01, based on two-tailed Student’s t-test. d & e, Representative IF images of three independent experiments (d) and quantification (e) of interphase cells treated with control siRNA or RTEL1 3’UTR siRNA for 48 hours and stained with S9.6 antibody (red) and FANCD2 antibody (green). A zoomed image for the area indicated by a white box is shown on the right. Error bars represent median ± s.e.m. (N = 3). ****p < 0.0001, based on two-tailed non-parametric Mann Whitney test. f & g, Representative IF images (f) and quantification (g) of MiDAS foci (labeled with EdU; red) in prometaphase cells treated as shown in panel a. DNA was stained with DAPI (blue). Error bars represent median ± s.e.m. (N = 3). ****p < 0.0001, based on two-tailed non-parametric Mann Whitney test. Scale bar is 10 µm. Data for graphs in panels c,e,g are available as Source data. Uncropped image for panel b is available as Source data.

Source data

Extended Data Fig. 2 RTEL1 is essential for MiDAS in telomerase positive cells under replication stress conditions.

a, Experimental workflow for the analysis of MiDAS in prometaphase HeLa or HCT116 cells following RTEL1 depletion and APH treatment. b & c, Representative images (b) and quantification (c) of RTEL1 expression following siRNA transfection with a control and RTEL1 siRNAs. Tubulin was used as a loading control. The quantification of relative protein levels was performed using Fiji/ImageJ software. Error bars indicate mean ± s.d. (N = 3). ****P < 0.0001, based on two-tailed Student’s t-test d & e, Representative IF images (d) and quantification (e) of MiDAS (indicated with EdU; red) in prometaphase cells. DNA was stained with DAPI (blue). Error bars represent median ± s.e.m. (N = 3). ****p < 0.0001, based on two-tailed non-parametric Mann Whitney test. Scale bar is 10 µm. Data for graphs in panels c,e are available as Source data. Uncropped image for panel b is available as Source data.

Source data

Extended Data Fig. 3 Following RTEL1 depletion, R-loops are accumulated in telomerase positive cells.

a, Experimental workflow for the analysis of R-loops in interphase HeLa or HCT116 cells following RTEL1 depletion. b & c, Representative IF images (b) and quantification (c) of interphase HeLa cells stained with S9.6 antibody specific for R-loops (red), or a nucleolin antibody (green). DNA was stained with DAPI (blue). Error bars represent median ± s.e.m. (N = 3). ****p < 0.0001, based on two-tailed non-parametric Mann Whitney test. d & e, Representative IF images (d) and quantification (e) of interphase HCT116 cells stained with S9.6 antibody specific for R-loops (red), or a nucleolin antibody (green). DNA was stained with DAPI (blue). The S9.6 signal intensity per nucleus was calculated by Fiji/ImageJ software and determined by subtracting the S9.6 staining with that from nucleoli (defined by nucleolin; green) in each nucleus. In each case, a specificity control for S9.6 staining is shown, in which fixed cells were treated with RNase H to eliminate R-loops. Error bars represent median ± s.e.m. (N = 3). ****p < 0.0001, based on two-tailed non-parametric Mann Whitney test. Scale bar is 10 µm. Data for graphs in panels c,e are available as Source data.

Source data

Extended Data Fig. 4 R-loops accumulate at specific regions of the genome following RTEL1 depletion.

a, Experimental workflow of cell synchronization and siRNA treatment for analyzing G1 phase U2OS cells. b, Representative IF images of G1 cells transfected with control siRNA or RTEL1 siRNA and stained with S9.6 antibody specific for R-loops (green) or a 53BP1 antibody (red). DNA was stained with DAPI (blue). Scale bar is 10 µm. c, Quantification of co-localization between 53BP1 bodies and R-loops in G1 cells from panel b. Error bars represent median ± s.e.m. (N = 3). ****p < 0.0001, based on two-tailed non-parametric Mann Whitney test. d, Experimental workflow for the quantitative analysis of R-loops enriched at different loci of the genome following RTEL1 depletion by DNA-RNA immunoprecipitation (DRIP) qPCR. e, Quantification of R-loops by DRIP qPCR at different genomic loci. The qPCR value was normalized against a region of the ZNF554 gene for each sample. Error bars indicate mean ± s.d. *p < 0.05; **p < 0.01, based on two-tailed Student’s t-test. (N for each locus: NRG3 = 8, FRA3B = 7, FRA16D = 4, FRA7B = 7, FRA10C = 3, 18 S = 5, 28 S = 3). Data for graphs in panels c,e are available as Source data.

Source data

Extended Data Fig. 5 RTEL1 depletion induces chromosome fragility at common fragile sites.

a, Experimental workflow for analyzing chromosome fragility in U2OS cells. b, Representative images of FISH signal on chromosomes from cells treated with either an siRNA control (siCon), an siRNA targeting RTEL1 (siRTEL1), or 0.4 μM APH. n, the number of loci examined. Loci of FRA16D (green) or FRA3B (red) were detected by FISH probes. DNA was stained with DAPI. c, Quantification of chromosome fragility at FRA16D or FRA3B. Error bars indicate mean ± s.d. (N = 3). *p < 0.05; **p < 0.01, based on two-tailed Student’s t-test. Scale bar is 2 µm. Data for graphs in panel c is available as Source data.

Source data

Extended Data Fig. 6 The analysis of un-bound protein and cell cycle profiles at different stages of interphase or mitosis following RTEL1 depletion and APH (0.4 µM) treatment.

The un-bound proteins was assessed by western blot analysis (a,c,e). GAPDH or Tubulin was used as loading controls. Histone H3 was used as a loading control for the chromatin-bound fraction. The cell cycle profiles (b,d,f) were output from Flowjo software and representative for two independent experiments. Uncropped images for panel a,c,e are available as Source data.

Extended Data Fig. 7 SLX4 depletion reduces the frequency of MiDAS in a U2OS cell line with inducible GFP-SLX4.

a, Experimental workflow of analyzing SLX4 in interphase cells or MiDAS in prometaphase cells by IF following SLX4 depletion. b, Representative images of a U2OS cell line with expression of GFP-SLX4 following the treatment with doxycycline (+Dox) to induce expression. GFP-SLX4 (green) or SLX4 (red) was detected by direct fluorescence (green; for GFP) or using antibodies against SLX4 respectively. c & d, Representative images of MiDAS (labeled with EdU; red) (c) and quantification (d) in prometaphase U2OS cells following SLX4 depletion by a 3’UTR siRNA or with inducible GFP-SLX4 expression. DNA was stained with DAPI (blue). Error bars represent median ± s.e.m. (N = 3). ****p < 0.0001, based on two-tailed non-parametric Mann Whitney test. Scale bar is 10 µm. Data for graphs in panel d is available as Source data.

Source data

Extended Data Fig. 8 SETX does not play a role in MiDAS.

a, Experimental workflow for analyzing prometaphase U2OS cells by IF following SETX depletion and APH treatment. b, Representative western blotting analysis of SETX following siRNA treatment. Actin was used as a loading control. c, d, Representative images of MiDAS (labeled with EdU; red) (c) and quantification (d) in prometaphase U2OS cells following SETX depletion. DNA was stained with DAPI (blue). Error bars represent median ± s.e.m. (N = 3). ns: p > 0.05, based on two-tailed non-parametric Mann Whitney test. Data for graphs in panel d is available as Source data. Uncropped image for panel b is available as Source data.

Source data

Extended Data Fig. 9 RNase H1 does not play a role in MiDAS.

a, Experimental workflow of analyzing interphase or prometaphase U2OS cells by IF following APH treatment and RNase H1 depletion alone or combined with RTEL1 depletion. b & c, Representative IF images (b) and quantification (c) of interphase cells stained with S9.6 antibody (red) and FANCD2 antibody (green). A zoomed image for the area indicated by a white box is shown on the right. Error bars represent median ± s.e.m. (N = 3). ****p < 0.0001, based on two-tailed non-parametric Mann Whitney test. d, Western blotting analysis of RNase H1 or RTEL1 following siRNA treatment. Actin was used as a loading control. e & f, Representative images of MiDAS (labeled with EdU; red) (e) and quantification (f) in prometaphase cells following RNaseH1 depletion alone or combined with RTEL1 depletion. Scale bar, 10 μm.DNA was stained with DAPI (blue). Error bars represent median ± s.e.m. (N = 3). ****p < 0.0001, based on two-tailed non-parametric Mann Whitney test. Data for graphs in panels c,f are available as Source data. Uncropped image for panel d is available as Source data.

Source data

Supplementary information

Supplementary Information

A cover page, reagents table, note, Figs. 1–3 and legends.

Reporting Summary

Supplementary Table 1

An Excel file of the peaks identified in DRIP-seq sample 7 (siRNA-control) and sample 9 (siRNA-RTEL1), excluding the peaks not affected by RNase H treatment. A P value was assigned to each peak, to assess whether it was differentially enriched in sample 9 against sample 7. In addition, these peaks are annotated with their genomic location and other features, including CFSs, G4 motifs, transcribed genes, COSMIC cancer mutations, COSMIC cancer gene expression and promoters.

Supplementary Table 2

An Excel file of the peaks identified in DRIP-seq sample 7 (siRNA-control) and sample 11 (siRNA-control+APH), excluding the peaks not affected by RNaseH treatment. A P value was assigned to each peak to assess whether it was differentially enriched in sample 11 against sample 7. In addition, these peaks have been annotated with their genomic location and other features, including CFSs, G4 motifs, transcribed genes, COSMIC cancer mutations, COSMIC cancer gene expression and promoters.

Supplementary Table 3

An Excel file of all of the CFSs annotated by three published databases in 2010, 2012 and 2019, respectively (see Supplementary Note).

Supplementary Data

Source data for unprocessed Western Blots and gels.

Source data

Source Data Fig. 1

Statistical source data

Source Data Fig. 2

Statistical source data

Source Data Fig. 3

Statistical source data

Source Data Fig. 4

Statistical source data

Source Data Fig. 5

Statistical source data

Source Data Fig. 6

Statistical source data

Source Data Fig. 7

Statistical source data

Source Data Fig. 8

Statistical source data

Source Data Extended Data Fig. 1

Statistical source data

Source Data Extended Data Fig. 2

Statistical source data

Source Data Extended Data Fig. 3

Statistical source data

Source Data Extended Data Fig. 4

Statistical source data

Source Data Extended Data Fig. 5

Statistical source data

Source Data Extended Data Fig. 7

Statistical source data

Source Data Extended Data Fig. 8

Statistical source data

Source Data Extended Data Fig. 9

Statistical source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wu, W., Bhowmick, R., Vogel, I. et al. RTEL1 suppresses G-quadruplex-associated R-loops at difficult-to-replicate loci in the human genome. Nat Struct Mol Biol 27, 424–437 (2020). https://doi.org/10.1038/s41594-020-0408-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-020-0408-6

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer