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:

The RNF138 E3 ligase displaces Ku to promote DNA end resection and regulate DNA repair pathway choice

Abstract

DNA double-strand breaks (DSBs) are repaired mainly by non-homologous end joining or homologous recombination (HR). Cell cycle stage and DNA end resection are believed to regulate the commitment to HR repair. Here we identify RNF138 as a ubiquitin E3 ligase that regulates the HR pathway. RNF138 is recruited to DNA damage sites through zinc fingers that have a strong preference for DNA with 5′- or 3′-single-stranded overhangs. RNF138 stimulates DNA end resection and promotes ATR-dependent signalling and DSB repair by HR, thereby contributing to cell survival on exposure to DSB-inducing agents. Finally, we establish that RNF138-dependent Ku removal from DNA breaks is one mechanism whereby RNF138 can promote HR. These results establish RNF138 as an important regulator of DSB repair pathway choice.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Screening E3s for HR repair.
Figure 2: RNF138 is recruited to sites of DSBs through its ZNF domains.
Figure 3: RNF138 regulates Ku80 ubiquitylation in response to DNA damage.
Figure 4: ChIP of endogenous RAD51, NBS1, Ku80 and MRE11 on a unique DSB.
Figure 5: Purification and DNA-binding activity of RNF138.
Figure 6: RNF138 knockdown results in hypersensitivity to DSB-inducing agents and impairs ATR-dependent signalling.
Figure 7: RNF138 acts at the early step of the HR pathway and is upstream of CtIP.
Figure 8: RNF138 regulates DNA end resection.

Similar content being viewed by others

References

  1. Hartlerode, A. J. & Scully, R. Mechanisms of double-strand break repair in somatic mammalian cells. Biochem. J. 423, 157–168 (2009).

    Article  CAS  Google Scholar 

  2. Pardo, B., Gomez-Gonzalez, B. & Aguilera, A. DNA repair in mammalian cells: DNA double-strand break repair: how to fix a broken relationship. Cell. Mol. Life Sci. 66, 1039–1056 (2009).

    Article  CAS  Google Scholar 

  3. Lieber, M. R. The mechanism of human nonhomologous DNA end joining. J. Biol. Chem. 283, 1–5 (2008).

    Article  CAS  Google Scholar 

  4. San Filippo, J., Sung, P. & Klein, H. Mechanism of eukaryotic homologous recombination. Annu. Rev. Biochem. 77, 229–257 (2008).

    Article  CAS  Google Scholar 

  5. Chapman, J. R., Taylor, M. R. & Boulton, S. J. Playing the end game: DNA double-strand break repair pathway choice. Mol. Cell 47, 497–510 (2012).

    Article  CAS  Google Scholar 

  6. Rodrigue, A. et al. Interplay between human DNA repair proteins at a unique double-strand break in vivo. EMBO J. 25, 222–231 (2006).

    Article  CAS  Google Scholar 

  7. Wu, D., Topper, L. M. & Wilson, T. E. Recruitment and dissociation of nonhomologous end joining proteins at a DNA double-strand break in Saccharomyces cerevisiae. Genetics 178, 1237–1249 (2008).

    Article  CAS  Google Scholar 

  8. Shim, E. Y. et al. Saccharomyces cerevisiae Mre11/Rad50/Xrs2 and Ku proteins regulate association of Exo1 and Dna2 with DNA breaks. EMBO J. 29, 3370–3380 (2010).

    Article  CAS  Google Scholar 

  9. Symington, L. S. & Gautier, J. Double-strand break end resection and repair pathway choice. Annu. Rev. Genet. 45, 247–271 (2011).

    Article  CAS  Google Scholar 

  10. Jazayeri, A. et al. ATM- and cell cycle-dependent regulation of ATR in response to DNA double-strand breaks. Nat. Cell Biol. 8, 37–45 (2006).

    Article  CAS  Google Scholar 

  11. Aylon, Y., Liefshitz, B. & Kupiec, M. The CDK regulates repair of double-strand breaks by homologous recombination during the cell cycle. EMBO J. 23, 4868–4875 (2004).

    Article  CAS  Google Scholar 

  12. Branzei, D. & Foiani, M. Regulation of DNA repair throughout the cell cycle. Nat. Rev. Mol. Cell Biol. 9, 297–308 (2008).

    Article  CAS  Google Scholar 

  13. Delacote, F., Han, M., Stamato, T. D., Jasin, M. & Lopez, B. S. An xrcc4 defect or Wortmannin stimulates homologous recombination specifically induced by double-strand breaks in mammalian cells. Nucl. Acids Res. 30, 3454–3463 (2002).

    Article  CAS  Google Scholar 

  14. Li, W. et al. Genome-wide and functional annotation of human E3 ubiquitin ligases identifies MULAN, a mitochondrial E3 that regulates the organelle’s dynamics and signaling. PLoS ONE 3, e1487 (2008).

    Article  Google Scholar 

  15. Karanam, K., Kafri, R., Loewer, A. & Lahav, G. Quantitative live cell imaging reveals a gradual shift between DNA repair mechanisms and a maximal use of HR in mid S phase. Mol. Cell 47, 320–329 (2012).

    Article  CAS  Google Scholar 

  16. Essers, J. et al. Nuclear dynamics of RAD52 group homologous recombination proteins in response to DNA damage. EMBO J. 21, 2030–2037 (2002).

    Article  CAS  Google Scholar 

  17. Katyal, S. et al. Aberrant topoisomerase-1 DNA lesions are pathogenic in neurodegenerative genome instability syndromes. Nat. Neurosci. 17, 813–821 (2014).

    Article  CAS  Google Scholar 

  18. Giannini, A. L., Gao, Y. & Bijlmakers, M. J. T-cell regulator RNF125/TRAC-1 belongs to a novel family of ubiquitin ligases with zinc fingers and a ubiquitin-binding domain. Biochem. J. 410, 101–111 (2008).

    Article  CAS  Google Scholar 

  19. Matsuoka, S. et al. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 316, 1160–1166 (2007).

    Article  CAS  Google Scholar 

  20. Shanbhag, N. M., Rafalska-Metcalf, I. U., Balane-Bolivar, C., Janicki, S. M. & Greenberg, R. A. ATM-dependent chromatin changes silence transcription in cis to DNA double-strand breaks. Cell 141, 970–981 (2010).

    Article  CAS  Google Scholar 

  21. Loughlin, F. E. & Mackay, J. P. Zinc fingers are known as domains for binding DNA and RNA. Do they also mediate protein-protein interactions? IUBMB Life 58, 731–733 (2006).

    Article  CAS  Google Scholar 

  22. Ismail, I. H. et al. CBX4-mediated SUMO modification regulates BMI1 recruitment at sites of DNA damage. Nucl. Acids Res. 40, 5497–5510 (2012).

    Article  CAS  Google Scholar 

  23. Yamada, M. et al. NARF, an nemo-like kinase (NLK)-associated ring finger protein regulates the ubiquitylation and degradation of T cell factor/lymphoid enhancer factor (TCF/LEF). J. Biol. Chem. 281, 20749–20760 (2006).

    Article  CAS  Google Scholar 

  24. Ardley, H. C. & Robinson, P. A. E3 ubiquitin ligases. Essays Biochem. 41, 15–30 (2005).

    Article  CAS  Google Scholar 

  25. Feng, L. & Chen, J. The E3 ligase RNF8 regulates KU80 removal and NHEJ repair. Nat. Struct. Mol. Biol. 19, 201–206 (2012).

    Article  CAS  Google Scholar 

  26. Postow, L. et al. Ku80 removal from DNA through double strand break-induced ubiquitylation. J. Cell Biol. 182, 467–479 (2008).

    Article  CAS  Google Scholar 

  27. Cheng, Q. et al. Ku counteracts mobilization of PARP1 and MRN in chromatin damaged with DNA double-strand breaks. Nucl. Acids Res. 39, 9605–9619 (2011).

    Article  CAS  Google Scholar 

  28. Britton, S., Coates, J. & Jackson, S. P. A new method for high-resolution imaging of Ku foci to decipher mechanisms of DNA double-strand break repair. J. Cell Biol. 202, 579–595 (2013).

    Article  CAS  Google Scholar 

  29. Bernstein, K. A. & Rothstein, R. At loose ends: resecting a double-strand break. Cell 137, 807–810 (2009).

    Article  CAS  Google Scholar 

  30. Zhou, Y. X. et al. A novel gene RNF138 expressed in human gliomas and its function in the glioma cell line U251. Anal. Cell. Pathol. 35, 167–178 (2012).

    Article  CAS  Google Scholar 

  31. Cimprich, K. A. & Cortez, D. ATR: an essential regulator of genome integrity. Nat. Rev. Mol. Cell Biol. 9, 616–627 (2008).

    Article  CAS  Google Scholar 

  32. Raderschall, E., Golub, E. I. & Haaf, T. Nuclear foci of mammalian recombination proteins are located at single-stranded DNA regions formed after DNA damage. Proc. Natl Acad. Sci. USA 96, 1921–1926 (1999).

    Article  CAS  Google Scholar 

  33. Pierce, A. J., Hu, P., Han, M., Ellis, N. & Jasin, M. Ku DNA end-binding protein modulates homologous repair of double-strand breaks in mammalian cells. Genes Dev. 15, 3237–3242 (2001).

    Article  CAS  Google Scholar 

  34. Shao, Z. et al. Persistently bound Ku at DNA ends attenuates DNA end resection and homologous recombination. DNA Repair 11, 310–316 (2012).

    Article  CAS  Google Scholar 

  35. Frank-Vaillant, M. & Marcand, S. Transient stability of DNA ends allows nonhomologous end joining to precede homologous recombination. Mol. Cell 10, 1189–1199 (2002).

    Article  CAS  Google Scholar 

  36. Kim, J. S. et al. Independent and sequential recruitment of NHEJ and HR factors to DNA damage sites in mammalian cells. J. Cell Biol. 170, 341–347 (2005).

    Article  CAS  Google Scholar 

  37. Schmidt, C. K. et al. Systematic E2 screening reveals a UBE2D-RNF138-CtIP axis promoting DNA repair. Nat. Cell Biol. 17, 1458–1470 (2015).

    Article  CAS  Google Scholar 

  38. Xie, A., Kwok, A. & Scully, R. Role of mammalian Mre11 in classical and alternative nonhomologous end joining. Nat. Struct. Mol. Biol. 16, 814–818 (2009).

    Article  CAS  Google Scholar 

  39. Pierce, A. J., Johnson, R. D., Thompson, L. H. & Jasin, M. XRCC3 promotes homology-directed repair of DNA damage in mammalian cells. Genes Dev. 13, 2633–2638 (1999).

    Article  CAS  Google Scholar 

  40. Mendez, J. & Stillman, B. Chromatin association of human origin recognition complex, cdc6, and minichromosome maintenance proteins during the cell cycle: assembly of prereplication complexes in late mitosis. Mol. Cell. Biol. 20, 8602–8612 (2000).

    Article  CAS  Google Scholar 

  41. Maity, R. et al. GST-His purification: a two-step affinity purification protocol yielding full-length purified proteins. J. Vis. Exp. e50320 (2013).

  42. Ismail, I. H. et al. CBX4-mediated SUMO modification regulates BMI1 recruitment at sites of DNA damage. Nucl. Acids Res. 40, 5497–5510 (2012).

    Article  CAS  Google Scholar 

  43. Menard, L. & Poirier, G. G. Rapid assay of poly(ADP-ribose) glycohydrolase. Biochem. Cell Biol. 65, 668–673 (1987).

    Article  CAS  Google Scholar 

  44. Shevchenko, A., Wilm, M., Vorm, O. & Mann, M. Mass spectrometric sequencing of proteins silver-stained polyacrylamide gels. Anal. Chem. 68, 850–858 (1996).

    Article  CAS  Google Scholar 

  45. Havlis, J., Thomas, H., Sebela, M. & Shevchenko, A. Fast-response proteomics by accelerated in-gel digestion of proteins. Anal. Chem. 75, 1300–1306 (2003).

    Article  CAS  Google Scholar 

  46. Shilov, I. V. et al. The Paragon Algorithm, a next generation search engine that uses sequence temperature values and feature probabilities to identify peptides from tandem mass spectra. Mol. Cell Proteomics 6, 1638–1655 (2007).

    Article  CAS  Google Scholar 

  47. Keller, A., Nesvizhskii, A. I., Kolker, E. & Aebersold, R. Empirical statistical model to estimate the accuracy of peptide identifications made by MS/MS and database search. Anal. Chem. 74, 5383–5392 (2002).

    Article  CAS  Google Scholar 

  48. Nesvizhskii, A. I., Keller, A., Kolker, E. & Aebersold, R. A statistical model for identifying proteins by tandem mass spectrometry. Anal. Chem. 75, 4646–4658 (2003).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank J. Stark (Department of Radiation Biology, Beckman Research Institute of the City of Hope, USA), R. Baer (Department of Pathology and Cell Biology, Columbia University, USA), S. M. Janicki (The Wistar Institute, USA), O. Hammarsten (Department of Clinical Chemistry, Goteborg University, Sweden), R. Greenberg (Department of Cancer Biology, Abramson Family Cancer Research Institute, USA) and J. Chen (Department of Experimental Radiation Oncology, The University of Texas MD Anderson Cancer Center, USA) for providing valuable reagents. We thank Y. Coulombe, M.-C. Caron and W. El Shemy for technical help. We also thank the Cellular Imaging Facility at the Cross Cancer Institute for the use of microscopes and image analysis software. The Proteomics Platform of the Quebec Genomics Center provided Mass spectrometry analysis. M.J.H. is an Alberta Innovates Health Solutions Senior Scholar. I.H.I. and H.S. are recipients of the Alberta Cancer Foundation postdoctoral fellowships. M.-M.G. was a CIHR Vanier scholar and is supported by a FRQS scholarship; G.G.P. holds a Tier 1 Canada Chair in Proteomics and J.-Y.M. is a FRQS Chercheur National Investigator. This work was supported by grants from the Canadian Institute of Health Research (CIHR).

Author information

Authors and Affiliations

Authors

Contributions

I.H.I. conceived the study and carried out most of the experiments. J.-P.G. and G.G.P. performed the mass spectrometry and PAR-binding experiments. M.-M.G. carried out the DNA binding assays. Z.X. generated all DNA constructs and mutants in the study. D.M. performed some immunofluorescence experiments. H.S. generated the GFP–RAD52/RFP–H2B cells. I.H.I., J.-Y.M., G.G.P. and M.J.H. analysed the data, conceived the figures and wrote the manuscript. All authors discussed the work and manuscript.

Corresponding authors

Correspondence to Ismail Hassan Ismail or Michael J. Hendzel.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 The RNF138 amino acid sequence is conserved.

(A) Amino acid alignment of RNF138 from different organisms is shown. Alignments was performed using Uniprot online tools. Identical residues are highlighted in black, conserved hydrophobic resides in light grey, other conservative changes are shown in darker grey. Star symbols indicate the conserved residues. (B) Scatter plots of the flow cytometric analysis of HeLa-Fucci cells prior to and post-sorting. The target cell populations were sorted using the indicated gates. (C) Western blot analysis of extracts prepared from sorted cells. Whole-cell extracts were prepared from sorted Fucci cells and probed for Ku80, cyclin A (S/G2 marker), and tubulin (loading control) by immunoblotting.

Supplementary Figure 2 RNF138 knock down alters Ku80 retention at the sites of DNA damage.

(A) GFP-Ku80 expressing cells were either transfected with control siRNA (CTL siRNA) or one of two different RNF138 siRNA (RNF138 siRNA1 and RNF138 siRNA2) for 48 h. Recruitment of GFP-Ku80 to sites of laser-induced DNA damage after micro-irradiation in U2OS cells was monitored using time lapse microscopy. Representative images are shown. Right: The mean levels of Ku80 accumulation at sites of laser tracks was quantified using LSM software and plotted as indicated. Data show mean ± s.d.; n = 44 cells, data pooled across two independent experiments. (B) RNF138 does not target Ku80 for degradation. U2OS cells transfected with either control siRNA (CTL siRNA) or RNF138 siRNA were treated with cycloheximide (CHX, 40 μM alone or together with MG132 (MG132, 20 μM) for indicated periods. The degradation of endogenous Ku80 was analyzed by anti-Ku80 immunoblotting. (C) Efficiency of NHEJ repair in NHEJ reporter cells reconstituted with different siRNA resistant RNF138 deletion mutant constructs and transfected with control siRNA (CTL siRNA) or RNF138-specific siRNA (RNF138 siRNA). The percentage of GFP positive cells was measured using a flow cytometer. Data show the mean ± SD; n = 3 biologically independent experiments that were based on gating 20,000 in each experimental condition. Scale bars are 5 μm.

Supplementary Figure 3 RNF138 mediates Ku80 ubiquitylation in response to DNA damage in S/G2 phase of the cell cycle.

Fucci cells were either transfected with control siRNA (CTL siRNA) or RNF138 specific siRNA (RNF138 siRNA) and then sorted by flow cytometry. Cells were exposed or not exposed to DNA damage by CLM. Soluble (Sol.) and chromatin fraction (Chr.) were isolated and subjected to immunopreciptation as indicated. Images are representative of two independent experiments.

Supplementary Figure 4 Cell cycle synchronization of DR95 cells expressing Isce-I cutting site.

Cells were synchronized, DNA content analysis was done by flow cytometer using propidium iodide staining. Left: Representatives DNA content profiles were shown. Right: the percentage of cells at each phase of the cell cycle is represented in the form of a table. Images are representative of two independent experiments.

Supplementary Figure 5 RNF138 knock down inhibits Ku removal from the sites of DNA damage in S phase.

(a) Cells were treated with CPT (1 μM, for 1 h) in the presence or absence of different treatments: ATM inhibitor (10 μM, added 1 h before CPT treatment, Mirin (100 μM, added 1 h before CPT treatment) or transfected with two different RNF138 siRNA (RNF138 siRNA1 and RNF138 siRNA2) for 48 h before CPT (1 μM, for 1 h). Images were captured using Z-stacks, deconvolved using Hygens deconvolution software, Ku80 foci were quantified using MetaXpress Software and plotted on the bottom. Average Ku80 foci per cell were counted and plotted as indicated. Data show mean ± s.d.; n = 42 cells that were obtained from two biological replicates. Scale bars are 10 μm.

Supplementary Figure 6 Effect of RNF138 depletion on the recruitment of Mre11 to sites of DNA damage.

(A) U2OS cells were co-transfected with the indicated siRNAs and YFP-Mre11 plasmids for 48 h. Cells were then laser-microirradiated and the recruitment of YFP-Mre11 to the sites of DNA damage was monitored by time-lapse microscopy. Signal intensity of YFP–Mre11 at DNA damage sites relative to the unirradiated area was quantified. Representative pictures are shown. (B) Western blot analysis of cells to estimate the knock down efficiency of RNF138 are shown and served as a control. (C) The mean levels of YFP-Mre11 accumulation at sites of laser tracks was quantified using LSM software and plotted as indicated. Data show mean ± s.d.; n = 34 cells, data pooled across two biologically independent experiments. Scale bars are 10 μm.

Supplementary Figure 7 RNF138 acts downstream of Mre11 activity in the HR pathway.

(A) U2OS cells were co-transfected with the indicated siRNAs and GFP-RNF138 for 48 h. In the case of Mirin treatment, cells transfected with GFP-RNF138 were pre-treated with Mirin (100 μM, added 1 h before laser microirradiation). Cells were then laser-microirradiated and the recruitment of GFP-RNF138 to the sites of DNA damage was monitored by time-lapse microscopy. Representative pictures were shown on the left. Quantification of GFP-RNF138 signal on laser-induced DNA damage stripes were shown on the right. Data show mean ± s.d.; n = 38 cells, data pooled across two biologically independent experiments. (B) Western blot analysis of cells to estimate the knock down efficiency of Mre11 and Ku80 are shown and served as a control. (C) Cell cycle profiles of RNF138 depleted cells. U2OS cells were transfected with control siRNA (CTL siRNA) or one of two different RNF138 siRNA (RNF138 siRNA1 and RNF138 siRNA2) for 48 h. Cells were harvested and subjected to DNA content analysis using propidium iodide staining according to standard protocols. Cell cycle graphs are shown on the left. A table summarizing the cell cycle results are shown on the right. At least 20,000 cells were analyzed. Scale bars are 5 μm.

Supplementary Figure 8 RNF138 did not appreciably alter MDC1/53BP1 IRIF.

Immunostaining of U2OS cells transfected with either control (CTL) siRNA or one of two different RNF138 siRNAs (RNF138 siRNA1 and RNF138 siRNA2) for 48 h. Cells were then exposed to 2 Gy and allowed to recover for 1 h before being fixed and immunostained as indicated. Bottom: quantification of populations of cells with >15 MDC1 foci colocalized with 53BP1 were plotted by counting 100 cells per experiment. Data show mean ± s.d.; n = 200 cells, data pooled across two biologically independent experiments. (B) U2OS cells were co-transfected with either control (CTL siRNA) or one of two different RNF138 siRNA (RNF138 siRNA1 and RNF138 siRNA2) targeting RNF138 and siRNA resistant GFP-RNF138 for 48 h. Cells were harvested and nuclear extracts were prepared and immunoblotted as indicated. Scale bars are 5 μm.

Supplementary information

Supplementary Information

Supplementary Information (PDF 1456 kb)

Supplementary Table 1

Supplementary Information (XLSX 27 kb)

Supplementary Table 2

Supplementary Information (XLS 26 kb)

Supplementary Table 3

Supplementary Information (XLSX 37 kb)

Supplementary Table 4

Supplementary Information (XLSX 33 kb)

Supplementary Table 5

Supplementary Information (XLSX 38 kb)

Supplementary Table 6

Supplementary Information (XLSX 17 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ismail, I., Gagné, JP., Genois, MM. et al. The RNF138 E3 ligase displaces Ku to promote DNA end resection and regulate DNA repair pathway choice. Nat Cell Biol 17, 1446–1457 (2015). https://doi.org/10.1038/ncb3259

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ncb3259

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