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:

Neuronal enhancers are hotspots for DNA single-strand break repair

Abstract

Defects in DNA repair frequently lead to neurodevelopmental and neurodegenerative diseases, underscoring the particular importance of DNA repair in long-lived post-mitotic neurons1,2. The cellular genome is subjected to a constant barrage of endogenous DNA damage, but surprisingly little is known about the identity of the lesion(s) that accumulate in neurons and whether they accrue throughout the genome or at specific loci. Here we show that post-mitotic neurons accumulate unexpectedly high levels of DNA single-strand breaks (SSBs) at specific sites within the genome. Genome-wide mapping reveals that SSBs are located within enhancers at or near CpG dinucleotides and sites of DNA demethylation. These SSBs are repaired by PARP1 and XRCC1-dependent mechanisms. Notably, deficiencies in XRCC1-dependent short-patch repair increase DNA repair synthesis at neuronal enhancers, whereas defects in long-patch repair reduce synthesis. The high levels of SSB repair in neuronal enhancers are therefore likely to be sustained by both short-patch and long-patch processes. These data provide the first evidence of site- and cell-type-specific SSB repair, revealing unexpected levels of localized and continuous DNA breakage in neurons. In addition, they suggest an explanation for the neurodegenerative phenotypes that occur in patients with defective SSB repair.

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: SAR-seq peaks occur within enhancers and are associated with PARP activation.
Fig. 2: Single-strand breaks detected by S1 END-seq after incubation with ddN.
Fig. 3: Localized SSB repair in neurons comprises short-patch and long-patch sub-pathways.
Fig. 4: Localized SSB repair in neurons correlates with sites of oxidized 5-methylcytosine.

Similar content being viewed by others

Data availability

The SAR-seq, S1 END-seq, Hi-C, RNA-seq, Seal and ChIP–seq data have been deposited in the Gene Expression Omnibus (GEO) database under accession number GSE167259. iPS cell enhancers were identified using H3K4me1 ChIP–seq from an ENCODE iPS cell line (GSM2527632). Experimentally validated enhancers were from VISTA genome browser for enhancers: https://enhancer.lbl.gov/.

References

  1. Caldecott, K. W. Single-strand break repair and genetic disease. Nat. Rev. Genet. 9, 619–631 (2008).

    Article  CAS  PubMed  Google Scholar 

  2. McKinnon, P. J. Genome integrity and disease prevention in the nervous system. Genes Dev. 31, 1180–1194 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Tubbs, A. & Nussenzweig, A. Endogenous DNA damage as a source of genomic instability in cancer. Cell 168, 644–656 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Miller, M. R. & Chinault, D. N. The roles of DNA polymerases alpha, beta, and gamma in DNA repair synthesis induced in hamster and human cells by different DNA damaging agents. J. Biol. Chem. 257, 10204–10209 (1982).

    Article  CAS  PubMed  Google Scholar 

  5. Fernandopulle, M. S. et al. Transcription factor-mediated differentiation of human iPSCs into neurons. Curr. Protoc. Cell Biol. 79, e51 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  6. Wang, C. et al. Scalable production of iPSC-derived human neurons to identify tau-lowering compounds by high-content screening. Stem Cell Reports 9, 1221–1233 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. van der Raadt, J., van Gestel, S. H. C., Nadif Kasri, N. & Albers, C. A. ONECUT transcription factors induce neuronal characteristics and remodel chromatin accessibility. Nucleic Acids Res. 47, 5587–5602 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  10. Song, M. et al. Mapping cis-regulatory chromatin contacts in neural cells links neuropsychiatric disorder risk variants to target genes. Nat. Genet. 51, 1252–1262 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Visel, A., Minovitsky, S., Dubchak, I. & Pennacchio, L. A. VISTA Enhancer Browser—a database of tissue-specific human enhancers. Nucleic Acids Res. 35, D88–D92 (2007).

    Article  CAS  PubMed  Google Scholar 

  12. Gupte, R., Liu, Z. & Kraus, W. L. PARPs and ADP-ribosylation: recent advances linking molecular functions to biological outcomes. Genes Dev. 31, 101–126 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Hanzlikova, H. & Caldecott, K. W. Perspectives on PARPs in S phase. Trends Genet. 35, 412–422 (2019).

    Article  CAS  PubMed  Google Scholar 

  14. Gibson, B. A., Conrad, L. B., Huang, D. & Kraus, W. L. Generation and characterization of recombinant antibody-like ADP-ribose binding proteins. Biochemistry 56, 6305–6316 (2017).

    Article  CAS  PubMed  Google Scholar 

  15. Madabhushi, R. et al. Activity-induced DNA breaks govern the expression of neuronal early-response genes. Cell 161, 1592–1605 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Suberbielle, E. et al. Physiologic brain activity causes DNA double-strand breaks in neurons, with exacerbation by amyloid-β. Nat. Neurosci. 16, 613–621 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Canela, A. et al. Topoisomerase II-induced chromosome breakage and translocation is determined by chromosome architecture and transcriptional activity. Mol. Cell 75, 252–266.e8 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Gómez-Herreros, F. et al. TDP2 suppresses chromosomal translocations induced by DNA topoisomerase II during gene transcription. Nat. Commun. 8, 233 (2017).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  19. Canela, A. et al. DNA breaks and end resection measured genome-wide by end sequencing. Mol. Cell 63, 898–911 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Caldecott, K. W. XRCC1 protein; form and function. DNA Repair 81, 102664 (2019).

    Article  CAS  PubMed  Google Scholar 

  21. Hanzlikova, H., Gittens, W., Krejcikova, K., Zeng, Z. & Caldecott, K. W. Overlapping roles for PARP1 and PARP2 in the recruitment of endogenous XRCC1 and PNKP into oxidized chromatin. Nucleic Acids Res. 45, 2546–2557 (2017).

    CAS  PubMed  Google Scholar 

  22. Caldecott, K. W. DNA single-strand break repair. Exp. Cell Res. 329, 2–8 (2014).

    Article  CAS  PubMed  Google Scholar 

  23. Caldecott, K. W. Mammalian DNA base excision repair: dancing in the moonlight. DNA Repair 93, 102921 (2020).

    Article  CAS  PubMed  Google Scholar 

  24. Tian, R. et al. CRISPR interference-based platform for multimodal genetic screens in human iPSC-derived neurons. Neuron 104, 239–255.e12 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Beard, W. A., Horton, J. K., Prasad, R. & Wilson, S. H. Eukaryotic base excision repair: new approaches shine light on mechanism. Annu. Rev. Biochem. 88, 137–162 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. DiGiuseppe, J. A., Hunting, D. J. & Dresler, S. L. Aphidicolin-sensitive DNA repair synthesis in human fibroblasts damaged with bleomycin is distinct from UV-induced repair. Carcinogenesis 11, 1021–1026 (1990).

    Article  CAS  PubMed  Google Scholar 

  27. Poetsch, A. R. The genomics of oxidative DNA damage, repair, and resulting mutagenesis. Comput. Struct. Biotechnol. J. 18, 207–219 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Bansal, K., Yoshida, H., Benoist, C. & Mathis, D. The transcriptional regulator Aire binds to and activates super-enhancers. Nat. Immunol. 18, 263–273 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Puc, J. et al. Ligand-dependent enhancer activation regulated by topoisomerase-I activity. Cell 160, 367–380 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Kalasova, I. et al. Pathological mutations in PNKP trigger defects in DNA single-strand break repair but not DNA double-strand break repair. Nucleic Acids Res. 48, 6672–6684 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Whitehouse, C. J. et al. XRCC1 stimulates human polynucleotide kinase activity at damaged DNA termini and accelerates DNA single-strand break repair. Cell 104, 107–117 (2001).

    Article  CAS  PubMed  Google Scholar 

  32. Lio, C. J. et al. TET methylcytosine oxidases: new insights from a decade of research. J. Biosci. 45, 21 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Kriaucionis, S. & Heintz, N. The nuclear DNA base 5-hydroxymethylcytosine is present in Purkinje neurons and the brain. Science 324, 929–930 (2009).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  34. Steinacher, R. et al. SUMOylation coordinates BERosome assembly in active DNA demethylation during cell differentiation. EMBO J. 38, e99242 (2019).

    Article  PubMed  CAS  Google Scholar 

  35. Song, C. X. et al. Genome-wide profiling of 5-formylcytosine reveals its roles in epigenetic priming. Cell 153, 678–691 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Szulwach, K. E. et al. 5-hmC-mediated epigenetic dynamics during postnatal neurodevelopment and aging. Nat. Neurosci. 14, 1607–1616 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Hoch, N. C. et al. XRCC1 mutation is associated with PARP1 hyperactivation and cerebellar ataxia. Nature 541, 87–91 (2017).

    Article  ADS  CAS  PubMed  Google Scholar 

  38. Lodato, M. A. et al. Aging and neurodegeneration are associated with increased mutations in single human neurons. Science 359, 555–559 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  39. Ried, D. A. et al. Incorporation of a nucleoside analog maps genome repair sites in post-mitotic human neurons. Science (in the press) (2021).

  40. Watanabe, S. et al. MyoD gene suppression by Oct4 is required for reprogramming in myoblasts to produce induced pluripotent stem cells. Stem Cells 29, 505–516 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Akiyama, T. et al. Efficient differentiation of human pluripotent stem cells into skeletal muscle cells by combining RNA-based MYOD1-expression and POU5F1-silencing. Sci. Rep. 8, 1189 (2018).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  42. Selvaraj, S. et al. Screening identifies small molecules that enhance the maturation of human pluripotent stem cell-derived myotubes. eLife 8, e47970 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Pawlowski, M. et al. Inducible and deterministic forward programming of human pluripotent stem cells into neurons, skeletal myocytes, and oligodendrocytes. Stem Cell Reports 8, 803–812 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Gilbert, L. A. et al. Genome-scale CRISPR-mediated control of gene repression and activation. Cell 159, 647–661 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Farías, G. G., Britt, D. J. & Bonifacino, J. S. Imaging the polarized sorting of proteins from the Golgi complex in live neurons. Methods Mol. Biol. 1496, 13–30 (2016).

    Article  PubMed  CAS  Google Scholar 

  46. Kirwan, P., Jura, M. & Merkle, F. T. Generation and characterization of functional human hypothalamic neurons. Curr. Protoc. Neurosci. 81, 3.33.1–3.33.24 (2017).

    Article  Google Scholar 

  47. Wong, N., John, S., Nussenzweig, A. & Canela, A. END-seq: an unbiased, high-resolution, and genome-wide approach to map DNA double-strand breaks and resection in human cells. Methods Mol. Biol. 2153, 9–31 (2021).

    Article  CAS  PubMed  Google Scholar 

  48. Bredemeyer, A. L. et al. DNA double-strand breaks activate a multi-functional genetic program in developing lymphocytes. Nature 456, 819–823 (2008).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  49. Santos, M. A. et al. DNA-damage-induced differentiation of leukaemic cells as an anti-cancer barrier. Nature 514, 107–111 (2014).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  50. Gibson, D. G. et al. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat. Methods 6, 343–345 (2009).

    Article  CAS  PubMed  Google Scholar 

  51. Noordermeer, S. M. et al. The shieldin complex mediates 53BP1-dependent DNA repair. Nature 560, 117–121 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  52. Cui, X. L. et al. A human tissue map of 5-hydroxymethylcytosines exhibits tissue specificity through gene and enhancer modulation. Nat. Commun. 11, 6161 (2020).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  53. Dai, Q. & He, C. Syntheses of 5-formyl- and 5-carboxyl-dC containing DNA oligos as potential oxidation products of 5-hydroxymethylcytosine in DNA. Org. Lett. 13, 3446–3449 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Rao, S. S. et al. A 3D map of the human genome at kilobase resolution reveals principles of chromatin looping. Cell 159, 1665–1680 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Buenrostro, J. D., Giresi, P. G., Zaba, L. C., Chang, H. Y. & Greenleaf, W. J. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat. Methods 10, 1213–1218 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Langmead, B., Trapnell, C., Pop, M. & Salzberg, S. L. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol. 10, R25 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  57. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    CAS  PubMed  Google Scholar 

  59. Li, H. et al. The sequence alignment/map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  62. Zang, C. et al. A clustering approach for identification of enriched domains from histone modification ChIP-seq data. Bioinformatics 25, 1952–1958 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Amemiya, H. M., Kundaje, A. & Boyle, A. P. The ENCODE blacklist: identification of problematic regions of the genome. Sci. Rep. 9, 9354 (2019).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  64. Salmon-Divon, M., Dvinge, H., Tammoja, K. & Bertone, P. PeakAnalyzer: genome-wide annotation of chromatin binding and modification loci. BMC Bioinformatics 11, 415 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  65. Harrow, J. et al. GENCODE: the reference human genome annotation for The ENCODE Project. Genome Res. 22, 1760–1774 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Whyte, W. A. et al. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell 153, 307–319 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Huang, W., Sherman, B. T. & Lempicki, R. A. Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat. Protocols 4, 44–57 (2009).

    Article  CAS  Google Scholar 

  68. Machanick, P. & Bailey, T. L. MEME-ChIP: motif analysis of large DNA datasets. Bioinformatics 27, 1696–1697 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Durand, N. C. et al. Juicebox provides a visualization system for Hi-C contact maps with unlimited zoom. Cell Syst. 3, 99–101 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Kent, W. J., Zweig, A. S., Barber, G., Hinrichs, A. S. & Karolchik, D. BigWig and BigBed: enabling browsing of large distributed datasets. Bioinformatics 26, 2204–2207 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Kent, W. J. et al. The human genome browser at UCSC. Genome Res. 12, 996–1006 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Ramírez, F. et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 44, W160–W165 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

We thank A. Rao, S. Wilson, S. Ruiz, N. Lopez-Bigas, D. Wilson III, Y. Sun, Y. Pommier, M. Adamowicz, P. Meltzer, and K. Ge for discussions and reagents; R. Chari for constructing the CAS9D10A inducible nickase; and the CCR genomics core for help with sequencing. K.W.C. is supported by Programme Grants from the UK Medical Research Council (MR/P010121/1), Cancer Research-UK (C6563/A7322), is an ERC Advanced Investigator (SIDSCA 694996) and received a Royal Society Wolfson Research Merit Award. The M.E.W. laboratory is supported by the NINDS Intramural Research Program, the Chan Zuckerberg Initiative, and the Packard ALS Center. S.E.H. received funding from the BrightFocus Foundation. The A.N. laboratory is supported by the Intramural Research Program of the NIH, an Ellison Medical Foundation Senior Scholar in Aging Award (AG-SS-2633-11), Department of Defense Awards (W81XWH-16-1-599 and W81XWH-19-1-0652), an Alex’s Lemonade Stand Foundation Award, and an NIH Intramural FLEX Award.

Author information

Authors and Affiliations

Authors

Contributions

W.W., S.E.H., W.J.N., J.P., K.W.C., M.E.W. and A.N. conceived, designed and analysed the experiments; W.W. designed bioinformatics pipelines, performed data analysis and designed the figures; S.E.H., E.C. and M.P. prepared i3Neurons for experiments; S.E.H. and E.C. performed gene knockdown, i3Neuron treatments, dissociation and collections in i3Neurons; J.P. and D.W. performed SAR-seq; W.J.N., J.P. and D.W. performed ChIP-seq; W.J.N. and N.V.W. performed S1 END-seq. S.E.H., E.C. and W.J.N. performed immunofluorescence imaging; A.C. developed SAR-seq in the A.N. laboratory; W.J.N. developed S1 END-seq; S.E.H. and M.E.W. selected guides for gene knockdown; K.S. performed ATAC-seq, did cloning for knock-down experiments and qRT–PCR; D.W. performed SEAL experiments; J.C.-M. prepared iMuscle cells; R.D.P. prepared primary rat neurons; S.E.H. treated and collected iMuscle and primary rat neurons; H.-Y.S. and S.C. conducted sequencing experiments in i3Neurons that were informative to the study; R.P. made MCF10A cells with Cas9-nickase; D.Z. and C.C. assisted with experiments; S.P. performed Hi-C, and S.K.J. and R.C. analysed the Hi-C data; H.H., P.J.M. and C.C. provided insights; K.W.C., M.E.W. and A.N. wrote the paper with input from all co-authors; M.E.W. and A.N. supervised the study.

Corresponding authors

Correspondence to Keith W. Caldecott, Michael E. Ward or André Nussenzweig.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks Michael Weinfeld and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Extended data figures and tables

Extended Data Fig. 1 Discrete genomic loci in neurons are associated with ongoing DNA synthesis.

a, Schematic of SAR-seq (DNA synthesis associated with repair sequencing) methodology. Neurons grown in culture (1) are incubated with EdU to label sites of DNA repair synthesis (2). The incorporated genomic EdU is then conjugated to biotin via click chemistry (3), sheared by sonication to fragments of 150–200 bp and captured with streptavidin beads (4). Enriched DNA sequences are then PCR-amplified and subjected to next-generation sequencing (5). b, Genome browser screenshot displaying SAR-seq profiles as normalized read density (RPM) for i3Neurons. Three independent biological replicates are shown as well as input. Neurons were labelled with EdU for 18 h and collected on day 7 after induction of neuronal differentiation. All coordinates provided are from the hg19 reference genome for i3Neurons. c, Venn diagram showing the overlap of SAR-seq peaks in i3Neurons for three independent biological replicates. d, Scatter plots showing correlations of SAR-seq intensities (RPKM) between three replicates in i3Neurons. Pearson correlation coefficients and P values are indicated. e, Left, representative images of EdU–biotin staining (green) showing cell proliferation in iPS cells, but not in post-mitotic i3Neurons. i3Neurons were treated with EdU on day 3 or day 6 and fixed on day 7. iPS cells were treated with EdU for 24 h and fixed. Cells were counterstained with DAPI (blue). Different imaging conditions were used for iPS cells and i3Neurons in the representative images. Right, quantification of EdU-positive cells. Each dot represents the percentage of dividing cells in one image (iPS cell: n = 8 images, n = 397/410 cells EdU-positive; i3Neuron day 3: n = 8 images, n = 35/483 cells EdU-positive; i3Neuron day 7: n = 8 images, n = 0/523 cells EdU-positive). Data are mean ± s.e.m. and are representative of three independent experiments.

Extended Data Fig. 2 Genomic characteristics of SAR-seq peaks.

a, Genome browser screenshot showing SAR-seq in i3Neurons treated with hydroxyurea (HU, n = 1) or polymerase-α inhibitor (POLαi, n = 2). NT, non-treated. b, Scatter plots showing SAR-seq intensities (RPKM) for HU- (left) and POLαi- (right) treated compared to non-treated (NT) samples. c, Flow cytometry cell cycle profile of pre-B cells treated or non-treated (NT) with POLαi. Cells were pulsed with EdU for 30 min before being collected for flow cytometry. Cells were counterstained with DAPI (n = 1). For an example of the gating strategy used for flow cytometry in c and in Extended Data Fig. 9b, see Supplementary Fig. 1. d, Histogram of individual SAR-seq peak widths, revealing a mean peak width of 901 bp. e, Genome browser screenshot showing SAR-seq in i3Neurons collected after 1, 2, 4, 8, or 18 h of EdU incubation (n = 2 for each). f, Fraction of EdU labelling in i3Neurons (relative to maximum labelling at 18 h) as a function of time, calculated from the top 2,000 peaks. Red points and error bars represent mean and s.d. of the relative levels of EdU measured from experimental data. Black line represents the theoretical model after fitting, with k being the rate of EdU labelling.

Extended Data Fig. 3 SAR-seq enrichment at neuronal intragenic regions.

a, Genome browser screenshot of chromosome 7 showing lack of localized DNA synthesis in two independent biological replicates of SAR-seq performed in iMuscle cells (n = 2) incubated with EdU for 18 h compared to SAR-seq in i3Neurons as well as input DNA. b, Genome browser screenshot displaying SAR-seq peak at a representative AsiSI restriction enzyme site (tick mark). AsiSI expression was induced for 18 h (+Dox, n = 1) compared with non-treated (−Dox, n = 1) conditions in G0-arrested, Abelson virus-transformed mouse pre-B cells as described19. c, Distribution of SAR-seq peaks with respect to different genomic features compared to genome-wide distribution of the hg19 human reference genome. Promoters are defined as 1 kb upstream of transcription start sites. Distal intergenic represents promoter-excluded intergenic regions. d, Fold enrichment of SAR-seq peaks in intergenic regions, intragenic regions and expressed genes compared to 1,000 sets of randomly shuffled regions of the same sizes and chromosome distribution (one-sided Fisher’s exact test, ****P < 2.2 × 10−16; NS, not significant). e, Left, scatterplot showing correlation of SAR-seq intensity (RPKM) with transcript level of genes containing SAR-seq peaks measured by RNA-seq (FPKM; n = 3) in i3Neurons. Seventy-one per cent of SAR-seq peaks are at expressed genes (FPKM ≥ 0.1; red dashed line, FPKM = 0.1). Right, correlation of SAR-seq intensity with transcript levels of linked genes determined by pcHiC in i3Neurons (red dashed line, FPKM = 0.1). Spearman correlation coefficients and P values are indicated. f, Genome browser screenshot comparing SAR-seq with strand-specific SAR-seq (n = 3), which discriminates which strand is labelled with EdU in i3Neurons. Both strands show labelling in three biological replicates. Strand-specific SAR-seq reads are separated into positive (black) and negative (grey) strands.

Extended Data Fig. 4 Motif discovery of SAR peaks and comparison with ATAC–seq peaks.

a, Motif analysis for sequences within 500 bp of the summit of the top 5,000 SAR-seq peaks in i3Neurons. Top, the best motif discovered by the MEME suite (1,384 out of 5,000 sites have this motif). Middle, TOMTOM motif tool used to compare SAR-seq motif shown above with databases of known motifs. The transcription factor ONECUT1 was identified as the most similar motif and its consensus sequence is shown. P value for motif comparison with ONECUT1 motif is indicated. Bottom, position distribution of the best motif (top) within 500 bp of the SAR-seq peak summit. The best motif is centred on the SAR-seq peak summit. b, Venn diagram illustrating the overlap between H3K4me1 ChIP–seq, ATAC–seq and SAR-seq peaks in i3Neurons. The statistical significance of the overlaps between SAR-seq, H3K4me1 ChIP–seq and ATAC–seq peaks was determined using randomly shuffled data sets (n = 1,000) by one-sided Fisher’s exact test (the P value for overlap between H3K4me1 ChIP–seq and SAR-seq peaks is P < 2.2 × 10−16, and for ATAC–seq/SAR-seq peaks is P < 2.2 × 10−16). Fraction of different overlapping groups are labelled in red for SAR-seq peaks, green for ATAC–seq peaks and blue for H3K4me1 ChIP–seq peaks. c, Heat maps of SAR-seq and ATAC–seq signals within 1 kb of SAR-seq peak summits in i3Neurons, ordered by SAR-seq intensity. d, Scatter plot comparing widths of ATAC–seq peaks and SAR-seq peaks for the top 10,000 overlapping peaks in i3Neurons. Pearson correlation coefficient and P value are indicated. e, Distribution of SAR-seq and ATAC–seq peaks with respect to different genomic features compared to genome-wide distribution on the hg19 human reference genome. Promoters are defined as 1 kb upstream of transcription start sites and distal intergenic represents promoter-excluded intergenic regions. f, Fold enrichment of SAR-seq and ATAC–seq peaks located at enhancers (black) and promoters (grey) compared to 1,000 sets of randomly shuffled regions of the same sizes and chromosome distributions, respectively (one-sided Fisher’s exact test, ****P < 2.2 × 10−16, NS: P = 0.0783, not significant). g, Heat map of RNA Pol II (n = 1) ChIP–seq and SAR-seq in i3Neurons for 1 kb on either side of the transcription start site (TSS) in i3Neurons, ordered by Pol II ChIP–seq intensity.

Extended Data Fig. 5 The correlation between SAR-seq and chromatin features.

a, Heat maps of SAR-seq and ChIP–seq data for enhancer markers (H3K4me1, H3K27ac and MLL4), other chromatin markers at accessible regions (H3K4me3, H3K36me3 and CTCF) and chromatin silencing markers (H3K27me3 and H3K9me3) for 1 kb on either side of the SAR-seq peak summit in i3Neurons, ordered by SAR-seq intensity. b, Scatter plots showing the correlation between SAR-seq intensity and ATAC–seq, H3K4me1, H3K27ac and MLL4 ChIP–seq intensities (RPKM) for 1 kb on either side of the SAR-seq peak summits in i3Neurons. Spearman correlation coefficients and P values are indicated. c, Fold enrichment of SAR-seq peaks at ATAC–seq peaks, ChIP–seq peaks of enhancer-related marks (H3K4me1, H3K27ac and MLL4), additional chromatin marks at accessible regions (H3K4me3, H3K36me3 and CTCF) (all n = 1) and chromatin silencing marks (H3K27me3 and H3K9me3) (both n = 1) in i3Neurons. n = 1,000 randomly shuffled data sets were generated to test the significance using one-sided Fisher’s exact test (P < 2.2 × 10−16 for H3K4me1, H3K27ac, ATAC–seq and MLL4; P = 1.85 × 10−316 for H3K4me3; P = 0.00116 for CTCF; **P < 0.001, ****P < 0.00001, NS, not significant). d, Heat maps of SAR-seq and H3K4me1 ChIP–seq signal for 1 kb on either side of the SAR-seq peak summit in primary rat cortical neurons, ordered by SAR-seq intensity. e, Venn diagram showing the overlap between H3K4me1 and SAR-seq peaks in rat primary neurons. n = 1,000 randomly shuffled data sets were generated to test the significance using one-sided Fisher’s exact test: P < 2.2 × 10−16. f, Scatter plot showing the correlation between SAR-seq and H3K4me1 ChIP–seq intensities (RPKM) for 1 kb on either side of the SAR-seq peak summits in rat primary neurons. Spearman correlation coefficient and P values are indicated.

Extended Data Fig. 6 SAR-seq enrichment at neuronal enhancers.

a, Box plot showing Hi-C contacts (left, n = 4) and H3K27ac (right, n = 1) levels at enhancers with (red) or without (grey) SAR-seq peaks. Contacts were defined as Hi-C interactions between an H3K27ac+ enhancer (with and without SAR) and its closest promoter within the TAD domain. For comparative purposes, observed contacts were normalized to expected contacts. Centre line, median; box limits, 25th and 75th percentiles; whisker indicates 1.5× interquartile range. Statistical significance was determined using the one-sided Wilcoxon rank-sum test. P = 1.64 × 10−225 for HiC contacts and P < 2.2 × 10−16 for H3K27ac intensity (****P < 0.00001). b, Fold enrichment of SAR-seq peaks at in vivo-validated enhancers from the VISTA Enhancer Browser database that overlap with H3K4me1 ChIP–seq peaks (left, P = 1.42 × 10−53) or at promoter-interacting regions determined by pcHiC (right, P = 1.574 × 10−09). n = 1,000 randomly shuffled datasets were generated to test significance using one-sided Fisher’s exact test (**P < 0.001, ****P < 0.00001). c, d, Genome browser screenshots showing SAR-seq, H3K4me1 and H3K27ac ChIP–seq, and pcHiC and Hi-C profiles at representative enhancers (highlighted in orange) interacting with the CDK5RAP3 promoter (c) and the DRD2 promoter (d). Both enhancers have been validated to promote transcription of their respective genes using CRIPSR techniques in i3Neurons10. The CDK5RAP3 enhancer also overlaps with in vivo-validated enhancers from the VISTA Enhancer Browser database. In the Hi-C contact matrix (c, bottom) the intensity of each pixel represents the normalized number of contacts between a pair of loci. The maximum intensity is indicated at the lower left corner. e, Pie chart showing distribution of i3Neuron SAR-seq peaks in iPS cell-specific, i3Neuron-specific and shared iPS cell and i3Neuron enhancers. Approximately 56 million and 49 million single end reads were sequenced for the H3K4me1 ChIP–seq in iPS cell and i3Neurons, respectively, with approximately 100,000 peaks called in both cell types. f, Top biological processes enriched for genes containing the 2,000 most intense SAR-seq peaks determined by GO analysis. The x-axis represents the enrichment value as the logarithm of FDR. G, H3K27ac signal at enhancers in i3Neurons ranked by H3K27ac ChIP–seq intensity. Red dashed line indicates the inflection point of the H3K27ac signal used to identify super-enhancers (cutoff 1,000). Accordingly, 1,385 enhancers were defined as super-enhancers. h, Bar graph showing the fraction of super-enhancers (left) and conventional enhancers (right) that overlap with SAR-seq peaks. The super-enhancers in the i3Neurons were defined by H3K27ac ChIP–seq intensity in g.

Extended Data Fig. 7 PARP and XRCC1 recruited to DNA repair sites.

a, Quantification of PAR levels with (n = 422 cells) and without (n = 541 cells) MMS treatment (red line, mean). Each dot represents one cell. Statistical significance was determined using two sided Mann–Whitney test (****P < 0.0001). Data are representative of three independent experiments. b, Anti-ADP-ribose ChIP–seq signal (n = 1) at an AsiSI restriction enzyme cut site (tick mark) in Abelson virus-transformed mouse pre-B cells. Cells were arrested in G0, and AsiSI double-strand breaks were induced for 18 h before ChIP. ADP-ribose is enriched at cleaved AsiSI sites and is increased by 20 min treatment with PARGi before fixation (AsiSI + PARGi), which is indicative of the presence of PAR. c, Genome browser screenshot illustrating the overlap between SAR-seq (n = 3), PAR (n = 1) and XRCC1 (n = 1) ChIP–seq signals in i3Neurons. Cells for PAR ChIP–seq were incubated with PARGi for 20 min before fixation.

Extended Data Fig. 8 Mapping regions of DNA damage and repair in neurons.

a, Genome browser example of SAR-seq profiles in non-treated (NT) or etoposide (ETO)-treated (18 h, 50 μM) i3Neurons. Data are from two biological replicates. b, Heat maps for SAR-seq in nontreated (NT) or etoposide- (ETO) treated (18 h, 50 μM) i3Neurons at −2 kb to +5 kb from the transcription start sites (TSS), ordered by ETO SAR-seq intensity. c, Immunofluorescence staining of the DSB markers γH2AX (red) and 53BP1 (green) in non-treated or ETO-treated (1 h) i3Neurons. Data are representative of three independent experiments. d, Genome browser showing SAR-seq and END-seq profiles in non-treated i3Neurons. END-seq, which detects DSBs specifically19, does not detect any enriched signal (that is, above background) at SAR-seq peaks. END-seq signals are separated into positive (black) and negative (grey) strands. END-seq data are representative of two independent experiments. e, Heat maps of SAR-seq and XRCC1 ChIP–seq (n = 1) for 1 kb on either side of SAR-seq peak summits in cultured rat primary neurons, ordered by SAR-seq intensity. f, g, Scatter plots showing the correlation between SAR-seq and XRCC1 ChIP–seq intensities (RPKM) for 1 kb before and after SAR-seq peak summits in i3Neurons (f) and rat primary neurons (g). Spearman correlation coefficients and P values are indicated.

Extended Data Fig. 9 S1 END-seq mapping of SSBs.

a, Top, genome browser screenshot showing profiles of SAR-seq and S1 END-seq. Agarose plugs were incubated with or without the restriction enzyme Nt.BspQI before S1 treatment (n = 1). Bottom, expanded views of Nt.BspQI sites (tick mark) show detection of S1 END-seq upon Nt.BspQI treatment. S1 END-seq reads are separated into positive (black) and negative (grey) strands. b, Flow cytometry profile of G1-arrested MCF10A cells pulsed with EdU. MCF10A cells were treated with palbociclib (1 μM) for 48 h to arrest cells in G1. Doxycycline was added in the last 24 h to induce nickase expression. Data are representative of three independent experiments. For an example of the gating strategy used for flow cytometry in b and Extended Data Fig. 2c, see Supplementary Fig. 1. c, Genome browser screenshot showing S1 END-seq profiles at three Cas9 nickase targeting sites (tick marks: sgRNAs 1–3) in G1-arrested MCF10A cells treated with doxycycline (+Dox) to induce Cas9 expression (n = 1). S1 END-seq signals are separated into positive (black) and negative (grey) strands. d, Heat maps of SAR-seq and S1 END-seq signals in i3Neurons with ddN or without incubation (NT) for 1 kb on either side of the SAR-seq peak summits, ordered by SAR-seq intensity. e, Venn diagram showing the overlap between S1 END-seq peaks incubated with ddN and SAR-seq peaks in i3Neurons. n = 1,000 randomly shuffled datasets were generated to test the significance using one-sided Fisher’s exact test: P < 2.2 × 10−16. f, Distribution of the size of the gaps between positive- and negative-strand S1 END-seq peak summits in i3Neurons incubated with ddN. The median gap size is 25 bp (red dashed line). The positive-strand peak represents the right end and the negative-strand peak represents the left end of a detected DSB. g, Aggregate plot showing the distribution of CG dinucleotides (black) for 1 kb on either side of SAR-seq peak summits overlaid with SAR-seq signal (red). h, Aggregate plot showing the distribution of CG dinucleotides (black) for 1 kb on either side of SAR-seq peak summits or summits of ATAC–seq peaks (green) that are H3K4me1 positive but do not overlap with SAR-seq.

Extended Data Fig. 10 PARP, XRCC1 or POLB deficiency increases SAR.

a, Quantitative RT–PCR analysis showing the PARP1 mRNA transcript level in i3Neurons after CRISPRi knockdown (sgControl, control non-targeting sgRNA; sgPARP1, an sgRNA targeting PARP1), cultured in parallel with samples used for SAR-seq. P = 7.72 × 10−7 by unpaired two-tailed Student’s t-test; ****P < 0.00001 (n = 3). b, Genome browser screenshot displaying SAR-seq profiles from two biological replicates of i3Neurons treated with the PARP inhibitors olaparib, talazoparib, or veliparib, or using CRISPRi-mediated knockdown with sgControl or sgPARP1, in duplicates. NT, not treated. c, Quantitative RT–PCR analysis showing XRCC1 mRNA transcript level in i3Neurons after CRISPRi knockdown, cultured in parallel with samples used for SAR-seq. P = 1.88 × 10−5 by unpaired two-tailed Student’s t-test; ****P < 0.00001 (n = 3). d, Genome browser screenshots of SAR-seq profiles in i3Neurons expressing sgControl or sgXRCC1, in duplicate. e, Venn diagram showing the overlap of SAR-seq peaks between i3Neurons expressing sgControl and those expressing sgXRCC1. n = 1,000 random datasets were generated to test the significance of overlap (one-sided Fisher’s exact test: P < 2.2 × 10−16). f, Quantitative RT–PCR analysis showing POLB mRNA transcript levels in i3Neurons after CRISPRi knockdown, cultured in parallel with samples used for SAR-seq. P = 6.98 × 10−5 for sgPOLB1 and 5.82 × 10−5 for sgPOLB2 by unpaired two-tailed Student’s t-test; ****P < 0.00001 (n = 3). g, Western blot showing POLB protein levels in i3Neurons after CRISPRi knockdown, cultured in parallel with samples used for SAR-seq (n = 1). For gel source data, see Supplementary Fig. 2. h, Genome browser screenshots of SAR-seq profiles from two biological replicates of i3Neurons expressing sgControl or sgPOLB, in duplicate. i, Genome browser screenshots of SAR-seq profiles from two biological replicates of i3Neurons expressing sgControl or sgPOLB. Cells were pre-treated or not treated (NT) with 50 μM Aph for 24 h, and then also during incubation with EdU.

Extended Data Fig. 11 Localized SSB repair in neurons correlates with sites of oxidized 5-methylcytosine.

a, Genome browser screenshot of chromosome 7 showing SAR-seq profiles from two biological replicates and input DNA in i3Neurons without MMS treatment (NT, n = 2) or after treatment with 0.1 mg ml−1 MMS (n = 2) for the final 15 min of an 18-h incubation with EdU. After streptavidin pull-down and PCR amplification, total DNA was quantified (NT rep 1: 0.95 μg; NT rep 2: 1.7 μg; MMS rep 1: 3.8 μg; MMS rep 2: 4.5 μg). Stochastic DNA damage results in loss of DNA synthesis at recurrent sites. b, Quantitative RT–PCR analysis showing PNKP mRNA transcript level in i3Neurons after CRISPRi knockdown, cultured in parallel with samples used for SAR-seq. P = 0.00015 by unpaired two-tailed Student’s t-test; ***P < 0.0001 (n = 3). c, Bottom, heat maps of SAR-seq intensities for 1 kb on either side of SAR-seq peak summits for i3Neurons expressing sgControl (n = 2) or sgPNKP (n = 2). Top, aggregate plots of SAR-seq intensity. d, Bottom, heat map of SAR-seq intensity for 1 kb on either side of the transcription start site in i3Neurons, ordered by SAR-seq intensity. i3Neurons expressing sgControl or sgPNKP were either not treated (NT, n = 2) or treated with 25 μM camptothecin (CPT, n = 2) during incubation with EdU. Top, aggregate plots of SAR-seq intensity. e, Scatter plots showing correlations of intensities (RPKM) between SSBs (ddN S1 END-seq) and 5fC or 5hmC, respectively, for 1 kb on either side of SAR-seq peak summits for i3Neurons. Spearman correlation coefficients and P values are indicated. f, Scatter plots showing correlations of intensities (RPKM) between SAR-seq for i3Neurons expressing sgPOLB and ddN S1 END-seq, 5fC, or 5hmC for 1 kb on either side of SAR-seq peak summits. Spearman correlation coefficients and P values are indicated.

Supplementary information

Supplementary Information

This file contains Supplementary Figure 1 (Flow cytometry gating) and Supplementary Figure 2 (Source Data gels).

Reporting Summary

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wu, W., Hill, S.E., Nathan, W.J. et al. Neuronal enhancers are hotspots for DNA single-strand break repair. Nature 593, 440–444 (2021). https://doi.org/10.1038/s41586-021-03468-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-021-03468-5

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