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:

Mechanism of 53BP1 activity regulation by RNA-binding TIRR and a designer protein

Abstract

Dynamic protein interaction networks such as DNA double-strand break (DSB) signaling are modulated by post-translational modifications. The DNA repair factor 53BP1 is a rare example of a protein whose post-translational modification-binding function can be switched on and off. 53BP1 is recruited to DSBs by recognizing histone lysine methylation within chromatin, an activity directly inhibited by the 53BP1-binding protein TIRR. X-ray crystal structures of TIRR and a designer protein bound to 53BP1 now reveal a unique regulatory mechanism in which an intricate binding area centered on an essential TIRR arginine residue blocks the methylated-chromatin-binding surface of 53BP1. A 53BP1 separation-of-function mutation that abolishes TIRR-mediated regulation in cells renders 53BP1 hyperactive in response to DSBs, highlighting the key inhibitory function of TIRR. This 53BP1 inhibition is relieved by TIRR-interacting RNA molecules, providing proof-of-principle of RNA-triggered 53BP1 recruitment to DSBs.

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

Fig. 1: TIRR blocks 53BP1 binding to NCP-ubme by masking the histone-binding surface of 53BP1.
Fig. 2: TIRR and 53BP1 anchor sites at the TIRR–53BP1 interface.
Fig. 3: Structure-based mutations in TIRR and 53BP1 affect complex formation and regulation of 53BP1 DSB recruitment.
Fig. 4: A separation-of-function 53BP1 mutation leads to a hyperactive form of 53BP1.
Fig. 5: Assessing the hyperactivity of 53BP1 F1553R mutant in multiple contexts.
Fig. 6: An engineered RNA- and nucleotide-processing enzyme, NUDT16TI, binds tightly to 53BP1.
Fig. 7: Nucleotides and RNA molecules dissociate the NUDT16TI–53BP1 and TIRR–53BP1 complexes.

Similar content being viewed by others

References

  1. Botuyan, M. V. et al. Structural basis for the methylation state-specific recognition of histone H4-K20 by 53BP1 and Crb2 in DNA repair. Cell 127, 1361–1373 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Fradet-Turcotte, A. et al. 53BP1 is a reader of the DNA-damage-induced H2A Lys 15 ubiquitin mark. Nature 499, 50–54 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Wilson, M. D. et al. The structural basis of modified nucleosome recognition by 53BP1. Nature 536, 100–103 (2016).

    Article  CAS  PubMed  Google Scholar 

  4. Mattiroli, F. et al. RNF168 ubiquitinates K13-15 on H2A/H2AX to drive DNA damage signaling. Cell 150, 1182–1195 (2012).

    Article  CAS  PubMed  Google Scholar 

  5. Gatti, M. et al. A novel ubiquitin mark at the N-terminal tail of histone H2As targeted by RNF168 ubiquitin ligase. Cell Cycle 11, 2538–2544 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Panier, S. & Boulton, S. J. Double-strand break repair: 53BP1 comes into focus. Nat. Rev. Mol. Cell Biol. 15, 7–18 (2014).

    Article  CAS  PubMed  Google Scholar 

  7. Bunting, S. F. et al. 53BP1 inhibits homologous recombination in Brca1-deficient cells by blocking resection of DNA breaks. Cell 141, 243–254 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Bouwman, P. et al. 53BP1 loss rescues BRCA1 deficiency and is associated with triple-negative and BRCA-mutated breast cancers. Nat. Struct. Mol. Biol. 17, 688–695 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Escribano-Díaz, C. et al. A cell cycle-dependent regulatory circuit composed of 53BP1-RIF1 and BRCA1-CtIP controls DNA repair pathway choice. Mol. Cell 49, 872–883 (2013).

    Article  CAS  PubMed  Google Scholar 

  10. Feng, L., Fong, K. W., Wang, J., Wang, W. & Chen, J. RIF1 counteracts BRCA1-mediated end resection during DNA repair. J. Biol. Chem. 288, 11135–11143 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Canny, M. D. et al. Inhibition of 53BP1 favors homology-dependent DNA repair and increases CRISPR-Cas9 genome-editing efficiency. Nat. Biotechnol. 36, 95–102 (2018).

    Article  CAS  PubMed  Google Scholar 

  12. Manis, J. P. et al. 53BP1 links DNA damage-response pathways to immunoglobulin heavy chain class-switch recombination. Nat. Immunol. 5, 481–487 (2004).

    Article  CAS  PubMed  Google Scholar 

  13. Ward, I. M. et al. 53BP1 is required for class switch recombination. J. Cell Biol. 165, 459–464 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Drané, P. et al. TIRR regulates 53BP1 by masking its histone methyl-lysine binding function. Nature 543, 211–216 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Zhang, A., Peng, B., Huang, P., Chen, J. & Gong, Z. The p53-binding protein 1-Tudor-interacting repair regulator complex participates in the DNA damage response. J. Biol. Chem. 292, 6461–6467 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Taylor, M. J. & Peculis, B. A. Evolutionary conservation supports ancient origin for Nudt16, a nuclear-localized, RNA-binding, RNA-decapping enzyme. Nucleic Acids Res. 36, 6021–6034 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. He, C. et al. High-resolution mapping of RNA-binding regions in the nuclear proteome of embryonic stem cells. Mol. Cell 64, 416–430 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Francia, S. et al. Site-specific DICER and DROSHA RNA products control the DNA-damage response. Nature 488, 231–235 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Burger, K. et al. Nuclear phosphorylated Dicer processes double-stranded RNA in response to DNA damage. J. Cell Biol. 216, 2373–2389 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Michelini, F. et al. Damage-induced lncRNAs control the DNA damage response through interaction with DDRNAs at individual double-strand breaks. Nat. Cell Biol. 19, 1400–1411 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Hu, Q., Botuyan, M. V., Cui, G., Zhao, D. & Mer, G. Mechanisms of ubiquitin-nucleosome recognition and regulation of 53BP1 chromatin recruitment by RNF168/169 and RAD18. Mol. Cell 66, 473–487.e9 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Ward, I. et al. The tandem BRCT domain of 53BP1 is not required for its repair function. J. Biol. Chem. 281, 38472–38477 (2006).

    Article  CAS  PubMed  Google Scholar 

  23. Zgheib, O., Pataky, K., Brugger, J. & Halazonetis, T. D. An oligomerized 53BP1 Tudor domain suffices for recognition of DNA double-strand breaks. Mol. Cell Biol. 29, 1050–1058 (2009).

    Article  CAS  PubMed  Google Scholar 

  24. Tang, J. et al. Acetylation limits 53BP1 association with damaged chromatin to promote homologous recombination. Nat. Struct. Mol. Biol. 20, 317–325 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Tong, Q. et al. Structural plasticity of methyllysine recognition by the tandem Tudor domain of 53BP1. Structure 23, 312–321 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Kim, H. et al. Crystal structure of syndesmos and its interaction with Syndecan-4 proteoglycan. Biochem. Biophys. Res. Commun. 463, 762–767 (2015).

    Article  CAS  PubMed  Google Scholar 

  27. Trésaugues, L. et al. Structural basis for the specificity of human NUDT16 and its regulation by inosine monophosphate. PLoS One 10, e0131507 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Lu, G. et al. hNUDT16: a universal decapping enzyme for small nucleolar RNA and cytoplasmic mRNA. Protein Cell 2, 64–73 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Iyama, T., Abolhassani, N., Tsuchimoto, D., Nonaka, M. & Nakabeppu, Y. NUDT16 is a (deoxy)inosine diphosphatase, and its deficiency induces accumulation of single-strand breaks in nuclear DNA and growth arrest. Nucleic Acids Res. 38, 4834–4843 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Abolhassani, N. et al. NUDT16 and ITPA play a dual protective role in maintaining chromosome stability and cell growth by eliminating dIDP/IDP and dITP/ITP from nucleotide pools in mammals. Nucleic Acids Res. 38, 2891–2903 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Daniels, C. M., Thirawatananond, P., Ong, S. E., Gabelli, S. B. & Leung, A. K. Nudix hydrolases degrade protein-conjugated ADP-ribose. Sci. Rep. 5, 18271 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Palazzo, L. et al. Processing of protein ADP-ribosylation by Nudix hydrolases. Biochem. J. 468, 293–301 (2015).

    Article  CAS  PubMed  Google Scholar 

  33. Zong, D. et al. Ectopic expression of RNF168 and 53BP1 increases mutagenic but not physiological non-homologous end joining. Nucleic Acids Res. 43, 4950–4961 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Dimitrova, N., Chen, Y. C., Spector, D. L. & de Lange, T. 53BP1 promotes non-homologous end joining of telomeres by increasing chromatin mobility. Nature 456, 524–528 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Margalef, P. et al. Stabilization of reversed replication forks by telomerase drives telomere catastrophe. Cell 172, 439–453.e14 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Rai, R. et al. The function of classical and alternative non-homologous end-joining pathways in the fusion of dysfunctional telomeres. EMBO J. 29, 2598–2610 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Smogorzewska, A., Karlseder, J., Holtgreve-Grez, H., Jauch, A. & de Lange, T. DNA ligase IV-dependent NHEJ of deprotected mammalian telomeres in G1 and G2. Curr. Biol. 12, 1635–1644 (2002).

    Article  CAS  PubMed  Google Scholar 

  38. Bothmer, A. et al. Regulation of DNA end joining, resection, and immunoglobulin class switch recombination by 53BP1. Mol. Cell 42, 319–329 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Callen, E. et al. 53BP1 mediates productive and mutagenic DNA repair through distinct phosphoprotein interactions. Cell 153, 1266–1280 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Pryde, F. et al. 53BP1 exchanges slowly at the sites of DNA damage and appears to require RNA for its association with chromatin. J. Cell Sci. 118, 2043–2055 (2005).

    Article  CAS  PubMed  Google Scholar 

  41. Lee, J., Thompson, J. R., Botuyan, M. V. & Mer, G. Distinct binding modes specify the recognition of methylated histones H3K4 and H4K20 by JMJD2A-tudor. Nat. Struct. Mol. Biol. 15, 109–111 (2008).

    Article  CAS  PubMed  Google Scholar 

  42. Mallette, F. A. et al. RNF8- and RNF168-dependent degradation of KDM4A/JMJD2A triggers 53BP1 recruitment to DNA damage sites. EMBO J. 31, 1865–1878 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Acs, K. et al. The AAA-ATPase VCP/p97 promotes 53BP1 recruitment by removing L3MBTL1 from DNA double-strand breaks. Nat. Struct. Mol. Biol. 18, 1345–1350 (2011).

    Article  CAS  PubMed  Google Scholar 

  44. Poulsen, M., Lukas, C., Lukas, J., Bekker-Jensen, S. & Mailand, N. Human RNF169 is a negative regulator of the ubiquitin-dependent response to DNA double-strand breaks. J. Cell Biol. 197, 189–199 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Panier, S. et al. Tandem protein interaction modules organize the ubiquitin-dependent response to DNA double-strand breaks. Mol. Cell 47, 383–395 (2012).

    Article  CAS  PubMed  Google Scholar 

  46. Chen, J., Feng, W., Jiang, J., Deng, Y. & Huen, M. S. Ring finger protein RNF169 antagonizes the ubiquitin-dependent signaling cascade at sites of DNA damage. J. Biol. Chem. 287, 27715–27722 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Jacquet, K. et al. The TIP60 complex regulates bivalent chromatin recognition by 53BP1 through direct H4K20me binding and H2AK15 acetylation. Mol. Cell 62, 409–421 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Yang, Q., Gilmartin, G. M. & Doublié, S. Structural basis of UGUA recognition by the Nudix protein CFI(m)25 and implications for a regulatory role in mRNA 3′ processing. Proc. Natl Acad. Sci. USA 107, 10062–10067 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Orthwein, A. et al. Mitosis inhibits DNA double-strand break repair to guard against telomere fusions. Science 344, 189–193 (2014).

    Article  CAS  PubMed  Google Scholar 

  50. Lee, D. H. et al. Dephosphorylation enables the recruitment of 53BP1 to double-strand DNA breaks. Mol. Cell 54, 512–525 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Otwinowski, Z. & Minor, W. Processing of X-ray diffraction data collected in oscillation mode. Methods Enzymol. 276, 307–326 (1997).

    Article  CAS  PubMed  Google Scholar 

  52. Adams, P. D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D. Biol. Crystallogr. 66, 213–221 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D. Biol. Crystallogr. 60, 2126–2132 (2004).

    Article  CAS  PubMed  Google Scholar 

  54. Delano, W. F. The PyMOL Molecular Graphics System, version 1.3r1. (Schrodinger, LLC, New York, 2010).

    Google Scholar 

  55. Brown, P. H., Balbo, A. & Schuck, P. Using prior knowledge in the determination of macromolecular size-distributions by analytical ultracentrifugation. Biomacromolecules 8, 2011–2024 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Drané, P., Ouararhni, K., Depaux, A., Shuaib, M. & Hamiche, A. The death-associated protein DAXX is a novel histone chaperone involved in the replication-independent deposition of H3.3. Genes Dev. 24, 1253–1265 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Karlseder, J., Smogorzewska, A. & de Lange, T. Senescence induced by altered telomere state, not telomere loss. Science 295, 2446–2449 (2002).

    Article  CAS  PubMed  Google Scholar 

  58. Ward, I. M., Minn, K., van Deursen, J. & Chen, J. p53 Binding protein 53BP1 is required for DNA damage responses and tumor suppression in mice. Mol. Cell. Biol. 23, 2556–2563 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We are very grateful to R. Alkire, N. Duke, and J. Lazarz at Argonne National Laboratory for their outstanding assistance. X-ray diffraction data were collected at Argonne National Laboratory, Structural Biology Center (SBC) at the Advanced Photon Source. SBC is operated by UChicago Argonne, LLC, for the US Department of Energy, Office of Biological and Environmental Research, under contract DE-AC02-06CH11357. This research was supported by NIH grants R01 CA132878, R01 GM116829, and P50 CA136393 (Mayo Clinic Ovarian Cancer SPORE developmental project) to G.M.; and by NIH grants R01 CA208244 and R01CA142698, DoD grant W81XWH-15-0564/OC140632, a Leukemia and Lymphoma Society Scholar grant, and the Claudia Adams Barr Program in Innovative Basic Cancer Research to D.C. M.V.B. was supported by DoD grant W81XWH-16-1-0391 and a Liz Tilberis award from the Ovarian Cancer Research Fund Alliance. G.C. received a Fellowship Award from the Mayo Clinic Cancer Center Fraternal Order of Eagles Funds. J.R.C. and C.O. were supported by a Cancer Research UK Career Development Fellowship Grant (C52690/A19270).

Author information

Authors and Affiliations

Authors

Contributions

M.V.B., G.C., P.D., D.C., and G.M. conceived the study. M.V.B., G.C., P.D., C.O., A.D., S.C., M.E.B., N.P., J.R.T., B.B., D.Z., J.R.C., D.C., and G.M. performed the experiments and/or analyzed the data. G.M. wrote the manuscript with extensive input from M.V.B., G.C., P.D., and D.C. All authors edited the manuscript.

Corresponding authors

Correspondence to Dipanjan Chowdhury or Georges Mer.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Integrated supplementary information

Supplementary Figure 1 X-ray structures of TIRR–53BP1 and NUDT16TI–53BP1.

a, The four TIRR and two 53BP1-Tudor molecules in the asymmetric unit are shown. b, The two NUDT16TI and two 53BP1-Tudor molecules in the asymmetric unit are shown

Supplementary Figure 2 Conformations of TIRR 53BP1-binding loop and corresponding loop region in NUDT16.

Left: Amino acid sequence alignment of TIRR 53BP1-binding loop with corresponding region of NUDT16. Right: Structural overlay of the loop regions in TIRR (blue) and NUDT16 (gray). Highlighted are Pro104 and Arg105 in NUDT16 and key 53BP1-interacting residues Pro105 and Arg107 in TIRR. 53BP1 residues interacting with TIRR Pro105 and Arg107 are also shown.

Supplementary Figure 3 A flexible loop in TIRR (residues 101–107) is highly specific for 53BP1 interaction.

Silver-stained gel (a) and immunoblot (b) of TIRR-FH and TIRR-Loop-FH partner proteins purified from the soluble nuclear extract of U2OS cells for mass spectrometric analysis. TIRR-Loop-FH corresponds to TIRR-FH mutated to harbor the loop region of NUDT16 (see Supplementary Fig. 2). Mass spectrometry data are in Supplementary Table 1

Supplementary Figure 4 Class switch recombination in stimulated B cells.

Shown are representative flow cytometry plots of the data presented in Fig. 5h

Supplementary Figure 5 Comparison of the X-ray structures of TIRR–53BP1 and NUDT16TI–53BP1.

a, Overlay of the TIRR–53BP1 and NUDT16TI–53BP1 structures with TIRR shown in blue, NUDT16TI in light blue, and 53BP1 in orange. The C-terminal α-helices in TIRR homodimer (shown in gray) do not exist in NUDT16TI. b, Details of the TIRR–53BP1 and NUDT16TI–53BP1 binding interfaces illustrating the remarkable similarity between the two complexes. Same color-coding as in a.

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Botuyan, M.V., Cui, G., Drané, P. et al. Mechanism of 53BP1 activity regulation by RNA-binding TIRR and a designer protein. Nat Struct Mol Biol 25, 591–600 (2018). https://doi.org/10.1038/s41594-018-0083-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-018-0083-z

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing