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:

EXD2 promotes homologous recombination by facilitating DNA end resection

Abstract

Repair of DNA double-strand breaks (DSBs) by homologous recombination (HR) is critical for survival and genome stability of individual cells and organisms, but also contributes to the genetic diversity of species. A vital step in HR is MRN–CtIP-dependent end resection, which generates the 3′ single-stranded DNA overhangs required for the subsequent strand exchange reaction. Here, we identify EXD2 (also known as EXDL2) as an exonuclease essential for DSB resection and efficient HR. EXD2 is recruited to chromatin in a damage-dependent manner and confers resistance to DSB-inducing agents. EXD2 functionally interacts with the MRN complex to accelerate resection through its 3′–5′ exonuclease activity, which efficiently processes double-stranded DNA substrates containing nicks. Finally, we establish that EXD2 stimulates both short- and long-range DSB resection, and thus, together with MRE11, is required for efficient HR. This establishes a key role for EXD2 in controlling the initial steps of chromosomal break 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

Figure 1: EXD2 is a CtIP interactor and its depletion sensitizes cells to DNA damage.
Figure 2: EXD2 depletion impairs DNA end resection following DSB induction.
Figure 3: EXD2 promotes HR and suppresses genome instability.
Figure 4: EXD2 displays 3′–5′ exonuclease activity in vitro.
Figure 5: Nuclease activity of EXD2 is required for DSB repair in vivo.
Figure 6: EXD2 promotes resection through a common mechanism with MRE11.
Figure 7: EXD2 is required for efficient HR.

Similar content being viewed by others

References

  1. Jackson, S. P. & Bartek, J. The DNA-damage response in human biology and disease. Nature 461, 1071–1078 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Malkova, A. & Haber, J. E. Mutations arising during repair of chromosome breaks. Annu. Rev. Genet. 46, 455–473 (2012).

    Article  CAS  PubMed  Google Scholar 

  3. Krogh, B. O. & Symington, L. S. Recombination proteins in yeast. Annu. Rev. Genet. 38, 233–271 (2004).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  5. Huertas, P., Cortes-Ledesma, F., Sartori, A. A., Aguilera, A. & Jackson, S. P. CDK targets Sae2 to control DNA-end resection and homologous recombination. Nature 455, 689–692 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Mimitou, E. P. & Symington, L. S. DNA end resection–unraveling the tail. DNA Rep. 10, 344–348 (2011).

    Article  CAS  Google Scholar 

  7. Stracker, T. H. & Petrini, J. H. The MRE11 complex: starting from the ends. Nat. Rev. Mol. Cell Biol. 12, 90–103 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Ciccia, A. & Elledge, S. J. The DNA damage response: making it safe to play with knives. Mol. Cell 40, 179–204 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Mimitou, E. P. & Symington, L. S. Sae2, Exo1 and Sgs1 collaborate in DNA double-strand break processing. Nature 455, 770–774 (2008).

    Article  CAS  PubMed  Google Scholar 

  10. Zhu, Z., Chung, W. H., Shim, E. Y., Lee, S. E. & Ira, G. Sgs1 helicase and two nucleases Dna2 and Exo1 resect DNA double-strand break ends. Cell 134, 981–994 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Cejka, P. et al. DNA end resection by Dna2-Sgs1-RPA and its stimulation by Top3-Rmi1 and Mre11-Rad50-Xrs2. Nature 467, 112–116 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Niu, H. et al. Mechanism of the ATP-dependent DNA end-resection machinery from Saccharomyces cerevisiae. Nature 467, 108–111 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Nimonkar, A. V. et al. BLM–DNA2–RPA–MRN and EXO1–BLM–RPA–MRN constitute two DNA end resection machineries for human DNA break repair. Genes Dev. 25, 350–362 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Paull, T. T. Making the best of the loose ends: Mre11/Rad50 complexes and Sae2 promote DNA double-strand break resection. DNA Rep. 9, 1283–1291 (2010).

    Article  CAS  Google Scholar 

  15. Blackwood, J. K. et al. End-resection at DNA double-strand breaks in the three domains of life. Biochem. Soc. Trans. 41, 314–320 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Sartori, A. A. et al. Human CtIP promotes DNA end resection. Nature 450, 509–514 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Nicolette, M. L. et al. Mre11-Rad50-Xrs2 and Sae2 promote 5′ strand resection of DNA double-strand breaks. Nat. Struct. Mol. Biol. 17, 1478–1485 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Cannavo, E. & Cejka, P. Sae2 promotes dsDNA endonuclease activity within Mre11-Rad50-Xrs2 to resect DNA breaks. Nature 514, 122–125 (2014).

    Article  CAS  PubMed  Google Scholar 

  19. Esashi, F., Galkin, V. E., Yu, X., Egelman, E. H. & West, S. C. Stabilization of RAD51 nucleoprotein filaments by the C-terminal region of BRCA2. Nat. Struct. Mol. Biol. 14, 468–474 (2007).

    Article  CAS  PubMed  Google Scholar 

  20. Paull, T. T. & Gellert, M. The 3′ to 5′ exonuclease activity of Mre 11 facilitates repair of DNA double-strand breaks. Mol. Cell 1, 969–979 (1998).

    Article  CAS  PubMed  Google Scholar 

  21. Garcia, V., Phelps, S. E., Gray, S. & Neale, M. J. Bidirectional resection of DNA double-strand breaks by Mre11 and Exo1. Nature 479, 241–244 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Jazayeri, A., Balestrini, A., Garner, E., Haber, J. E. & Costanzo, V. Mre11-Rad50-Nbs1-dependent processing of DNA breaks generates oligonucleotides that stimulate ATM activity. EMBO J. 27, 1953–1962 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Shibata, A. et al. DNA double-strand break repair pathway choice is directed by distinct MRE11 nuclease activities. Mol. Cell 53, 7–18 (2014).

    Article  CAS  PubMed  Google Scholar 

  24. Smogorzewska, A. et al. A genetic screen identifies FAN1, a Fanconi anemia-associated nuclease necessary for DNA interstrand crosslink repair. Mol. Cell 39, 36–47 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Limbo, O. et al. Ctp1 is a cell-cycle-regulated protein that functions with Mre11 complex to control double-strand break repair by homologous recombination. Mol. Cell 28, 134–146 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Yun, M. H. & Hiom, K. CtIP-BRCA1 modulates the choice of DNA double-strand-break repair pathway throughout the cell cycle. Nature 459, 460–463 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Kousholt, A. N. et al. CtIP-dependent DNA resection is required for DNA damage checkpoint maintenance but not initiation. J. Cell Biol. 197, 869–876 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Eggler, A. L., Inman, R. B. & Cox, M. M. The Rad51-dependent pairing of long DNA substrates is stabilized by replication protein A. J. Biol. Chem. 277, 39280–39288 (2002).

    Article  CAS  PubMed  Google Scholar 

  29. Sung, P., Krejci, L., Van Komen, S. & Sehorn, M. G. Rad51 recombinase and recombination mediators. J. Biol. Chem. 278, 42729–42732 (2003).

    Article  CAS  PubMed  Google Scholar 

  30. Sung, P. & Robberson, D. L. DNA strand exchange mediated by a RAD51-ssDNA nucleoprotein filament with polarity opposite to that of RecA. Cell 82, 453–461 (1995).

    Article  CAS  PubMed  Google Scholar 

  31. Baumann, P., Benson, F. E. & West, S. C. Human Rad51 protein promotes ATP-dependent homologous pairing and strand transfer reactions in vitro. Cell 87, 757–766 (1996).

    Article  CAS  PubMed  Google Scholar 

  32. Filippone, S. et al. On the mechanism of the thermal retrocycloaddition of pyrrolidinofullerenes (retro-Prato reaction). Chemistry 14, 5198–5206 (2008).

    Article  CAS  PubMed  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  PubMed  PubMed Central  Google Scholar 

  34. McCabe, N. et al. Deficiency in the repair of DNA damage by homologous recombination and sensitivity to poly(ADP-ribose) polymerase inhibition. Cancer Res. 66, 8109–8115 (2006).

    Article  CAS  PubMed  Google Scholar 

  35. Buis, J. et al. Mre11 nuclease activity has essential roles in DNA repair and genomic stability distinct from ATM activation. Cell 135, 85–96 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Choi, J. M. et al. Probing the roles of active site residues in the 3′–5′ exonuclease of the Werner syndrome protein. J. Biol. Chem. 282, 9941–9951 (2007).

    Article  CAS  PubMed  Google Scholar 

  37. Li, B., Reddy, S. & Comai, L. Sequence-specific processing of telomeric 3′ overhangs by the Werner syndrome protein exonuclease activity. Aging 1, 289–302 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Rass, E. et al. Role of Mre11 in chromosomal nonhomologous end joining in mammalian cells. Nat. Struct. Mol. Biol. 16, 819–824 (2009).

    Article  CAS  PubMed  Google Scholar 

  39. Huertas, P. DNA resection in eukaryotes: deciding how to fix the break. Nat. Struct. Mol. Biol. 17, 11–16 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Zhou, Y., Caron, P., Legube, G. & Paull, T. T. Quantitation of DNA double-strand break resection intermediates in human cells. Nucleic Acids Res. 42, e19 (2014).

    Article  CAS  PubMed  Google Scholar 

  41. Hsu, P. D., Lander, E. S. & Zhang, F. Development and applications of CRISPR-Cas9 for genome engineering. Cell 157, 1262–1278 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Shen, B. et al. Efficient genome modification by CRISPR-Cas9 nickase with minimal off-target effects. Nat. Methods 11, 399–402 (2014).

    Article  CAS  PubMed  Google Scholar 

  43. Suzuki, K., Yamauchi, M., Oka, Y., Suzuki, M. & Yamashita, S. Creating localized DNA double-strand breaks with microirradiation. Nat. Protoc. 6, 134–139 (2011).

    Article  CAS  PubMed  Google Scholar 

  44. Gravel, S., Chapman, J. R., Magill, C. & Jackson, S. P. DNA helicases Sgs1 and BLM promote DNA double-strand break resection. Genes Dev. 22, 2767–2772 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Rubnitz, J. & Subramani, S. The minimum amount of homology required for homologous recombination in mammalian cells. Mol. Cell. Biol. 4, 2253–2258 (1984).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Sugawara, N. & Haber, J. E. Characterization of double-strand break-induced recombination: homology requirements and single-stranded DNA formation. Mol. Cell. Biol. 12, 563–575 (1992).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Iacovoni, J. S. et al. High-resolution profiling of γH2AX around DNA double strand breaks in the mammalian genome. EMBO J. 29, 1446–1457 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Cong, L. et al. Multiplex genome engineering using CRISPR/Cas systems. Science 339, 819–823 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Savitsky, P. et al. High-throughput production of human proteins for crystallization: the SGC experience. J. Struct. Biol. 172, 3–13 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Tuschl, T. Cotransfection of luciferase reporter plasmids with siRNA duplexes. CSH Protoc. (2006)10.1101/pdb.prot4342.

  51. Blackford, A. N. et al. TopBP1 interacts with BLM to maintain Genome stability but is dispensable for preventing BLM degradation. Mol. Cell 57, 1133–1141 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Bhaskara, V. et al. Rad50 adenylate kinase activity regulates DNA tethering by Mre11/Rad50 complexes. Mol. Cell 25, 647–661 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Sengerova, B. et al. Characterization of the human SNM1A and SNM1B/Apollo DNA repair exonucleases. J. Biol. Chem. 287, 26254–26267 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Zhu, L. & Halligan, B. D. V(D)J recombinational signal sequence DNA binding activities expressed by fetal bovine thymus. Veter. Immunol. Immunopathol. 71, 277–289 (1999).

    Article  CAS  Google Scholar 

  55. Hsu, P. D. et al. DNA targeting specificity of RNA-guided Cas9 nucleases. Nat. Biotechnol. 31, 827–832 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. 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  PubMed  PubMed Central  Google Scholar 

  57. Wang, Q. et al. Rad17 recruits the MRE11-RAD50-NBS1 complex to regulate the cellular response to DNA double-strand breaks. EMBO J. 33, 862–877 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank S. Jackson, M. de Bruijn, J. Riepsaame, V. Macaulay, G. Legube, G. Stewart, F. Esashi, C. Green and R. Chapman for cell lines, plasmids and antibodies; J. A. Tainer for the provision of MRE11 exo- and endonuclease inhibitors; the Mass Spectrometry Laboratory (IBB PAS, Warsaw, Poland) for their work on analyses of GFP–CtIP IP experiments; J. A. Newman for advice on protein purification and D. Waithe for help with image analysis. We also thank G. Ira for critical reading of the manuscript. R.B., J.N. and W.N. are funded by Worldwide Cancer Research and MRC Senior Non-Clinical Fellowships awarded to W.N., P.J.M. and O.G. are funded by an MRC Grant, H.T.B. is funded by a Cancer Research UK Studentship and R.A.D. and T.T.P. are funded by the Cancer Research and Prevention Institute of Texas (CPRIT) grant RP110465-P4.

Author information

Authors and Affiliations

Authors

Contributions

R.B. and J.N. carried out the majority of experimental work with contributions from W.N.; H.T.B., P.J.M. and O.G. contributed to the purification and analysis of the biochemical activities of EXD2. R.A.D. and T.T.P. purified the MRN complex. W.N. conceived the project and wrote the manuscript with editing contributions from R.B., J.N., T.T.P. and P.J.M.

Corresponding author

Correspondence to Wojciech Niedzwiedz.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 5

Cartoon illustrating the phylogenetic tree for the EXD2 gene.

Supplementary Figure 6 Additional characterisation of resection defect in EXD2-depleted cells.

(a,b) Western blotting confirming depletion of endogenous EXD2 72 h post-transfection with either control siRNA (siControl) or siRNAs targeting EXD2 (siEXD2-1 and 2). α-Tubulin acts as a loading control. These experiments were carried out three times independently. (c) RPA foci in U2OS cells 72 h post-transfection with control siRNA (siControl) or an siRNA oligo targeting EXD2 (siEXD2). Cells were either untreated or were exposed to 8 Gy IR and left to recover for 1 h before fixation and then stained for RPA and DAPI as indicated. Scale bar = 20 μm. (d) Quantification of the percentage of cells treated as in (c), exhibiting greater than 15 RPA foci per nucleus. n = 300 cells (siControl untreated), 390 cells (siEXD2 untreated), 355 cells (siControl 8Gy IR) and 403 cells (siEXD2 8gy IR) respectively, pooled from three independent experiments. Bars represent mean values ± s.e.m. The Chi-square test was used to determine statistical significance. (e) Western blotting of various DDR proteins in U2OS cells 72 h post-transfection with control siRNA (siControl) or an siRNA oligo targeting EXD2 (siEXD2) with cells treated with 8 Gy IR. Samples were acquired at the indicated time points post-IR treatment. Chk2-p T68 acts as a control for ATM activation, RPA2 pS4/S8 acts to indicate resection efficiency, γH2AX serves to indicate DSB induction with RPA and histone H3 serving as loading controls. This experiment was carried out two times independently. (f) Quantification of the percentage of G1, S or G2/M U2OS cells as analysed by propidium iodide staining and FACS analysis. Cells were analysed 72 h post-transfection with control siRNA (siControl) or siRNA targeting EXD2 (siEXD2) (mean values ± s.e.m., n = 5 independent experiments).

Supplementary Figure 7 Additional characterisation of purified EXD2.

(a) Alignment of the partial amino acid sequences of EXD2 proteins from various vertebrate organisms (human, mouse, chicken and xenopus are depicted) with a partial amino acid sequence of the exonuclease domain of human Werner helicase protein (WRN). Highly conserved residues shown to mediate the exonuclease activity of WRN protein which are also conserved in human EXD2 and its vertebrate homologues are indicated with red arrows. (b) Coomassie stained SDS-PAGE gel of EXD2 WT or D108A E110A mutant protein ectopically expressed in E. coli and purified to homogeneity for use in in vitro biochemistry experiments. (c) InstantBlue stained SDS-PAGE gel of the truncated EXD2 proteins (WT and D108A E110A mutant) used in this study. (d) Mass spectrum of truncated EXD2 (K76–V564) WT protein confirming sample purity. (e) Mass spectrum of truncated EXD2 (K76–V564) D108A E110A protein confirming sample purity. (f) 5′ radiolabeled ssDNA or dsDNA substrate (10 nM molecules) was incubated for indicated amounts of time with EXD2 WT (K76–V564) or EXD2 (K76–V564) D108A E110A (EXD2 mut) protein (100 nM). Samples were resolved on a 20% TBE-Urea polyacrylamide gel and visualised by phosphorimaging. This experiment was carried out two times independently.

Supplementary Figure 8 Western blots confirming EXD2 knockdown efficiency and additional characterisation of EXD2 in vivo and in vitro.

(a) Western blotting determining the relative levels of expression of Flag–HA–EXD2 WT or D108A E110A mutant proteins in U2OS cells stably expressing these fusion proteins. Two independent clones for each construct are shown. Cells were transfected with control siRNA (siControl) or siRNA targeting endogenous EXD2 (siEXD2 3′UTR) as indicated and collected for western blotting. MCM2 serves as a loading control. This experiment was carried out three times independently. (b) Quantification of the frequency of RAD51 focus-positive U2OS cells 72 h post-transfection with the indicated siRNA. Cells were either untreated or exposed to 8 Gy IR and left to recover for 6 h before fixation. Cells were stained with DAPI and RAD51 as indicated. The percentage of cells exhibiting RAD51 foci was quantified. n = 366 cells (siControl 0 min), 386 cells (siEXD2 0 min), 337 cells (siMRE11 0 min), 315 cells (siEXD2/siMRE11, 0 min), n = 308 cells (siControl 360 min), 321 cells (siEXD2 360 min), 337 cells (siMRE11 360 min), 374 cells (siEXD2/siMRE11, 346 min), respectively, pooled from three independent experiments. Bars represent mean values ± s.e.m. The Chi-square test was used to determine statistical significance. (c) Western blotting confirming the depletion of EXD2 and MRE11 in U2OS cells 72 h post-transfection with control siRNA (siControl) or siRNA targeting EXD2 or MRE11 as indicated. MCM2 serves as a loading control. This experiment was carried out three times independently. (d,e) 5′ radiolabeled ssDNA or dsDNA 50-mer substrates (1 nM molecules) (d) or 5′ radiolabeled dsDNA substrates (1 nM molecules) containing a nick or 1 nucleotide gap (e) were incubated for the indicated amounts of time with EXD2 WT protein (70 nM). Samples were resolved on a 20% TBE-Urea polyacrylamide gel and visualised by phosphorimaging. These experiments were carried out once. (f,g) Western blotting confirming the depletion of EXD2 and MRE11 in ER-AsiSI U2OS cells (f) and DR-GFP U2OS cells (g) 72 h post-transfection with control siRNA (siControl) or siRNA targeting EXD2 or MRE11 as indicated. MCM2 serves as a loading control. These experiments were carried out three times independently.

Supplementary Figure 9 Generation and characterisation of EXD2 knockout cell lines.

(a) EXD2 knockout generation strategy using the CRISPR-Cas9 nickase. Schematic representation of the human EXD2 genomic locus with guide RNAs sequences highlighted in green and predicted cut sites marked by red arrows. (b) Representative images of RPA foci in HeLa control cells and in two independent clones of HeLa EXD2−/− cells treated with 1 μM CPT for 1 h. Scale bar = 5 μm. (c) Quantification of the percentage of HeLa or HeLa EXD2−/− cells treated as in b. exhibiting greater than 15 RPA foci per nucleus. Data from two independent HeLa EXD2−/− clones are represented. n = 617 cells (HeLa), 433 cells (HeLa EXD2−/− cl.1) and 429 (HeLa EXD2−/− cl.2) respectively, pooled from three independent experiments. Bars represent mean values ± s.e.m. The Chi-square test was used to determine statistical significance. (d) Western blot of HeLa EXD2−/− clones and parental cells treated with 1 μM CPT for 1 h. RPA2 pS4/S8 acts as an indicator of resection, MCM2 acts as a loading control. This experiment was carried out two times independently. (e) Survival of HeLa control cells and HeLa EXD2−/− cells treated with the indicated doses of CPT. Survival data from two independent EXD2−/− clones is depicted. Survival data represent mean ± s.e.m., (n = 3 independent experiments). (f) Western blot of HeLa EXD2−/− clones and parental cells probed with antibodies against MRE11 and CtIP. α-Tubulin acts as a loading control. This experiment was carried out three times independently.

Supplementary Figure 10 EXD2 is not required for CtIP or MRE11 recruitment to site of DNA damage.

(a) Immunofluorescence microscopy of U2OS cells stably expressing GFP-CtIP treated with control siRNA (siControl) or siRNA targeting EXD2 (siEXD2) following the induction of localized DSBs by microirradiation. DAPI serves as a marker for the cell nucleus and γH2AX serves as a marker for DSB induction. Scale bar = 20 μm. This experiment was carried out three times independently. (b) Western blotting confirming EXD2 knock down efficacy in samples from (a). α-Tubulin serves as a loading control. This experiment was carried out three times independently. (c) Immunofluorescence microscopy of WT HeLa and HeLa EXD2−/− clones stained for MRE11 using an antibody recognising the endogenous protein following the induction of localized DSBs by microirradiation. γH2AX serves as a marker for DSB induction. Scale bar = 20 μm. This experiment was carried out three times independently.

Supplementary Figure 11

Original uncropped images of western blots.

Supplementary Table 1 Antibodies.
Supplementary Table 2 siRNA sequences.
Supplementary Table 3 DNA oligonucleotides used in in vitro nuclease assays.
Supplementary Table 4 qPCR primer sequences used in ER-AsiSI resection assay.

Supplementary information

Supplementary Information

Supplementary Information (PDF 0 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Broderick, R., Nieminuszczy, J., Baddock, H. et al. EXD2 promotes homologous recombination by facilitating DNA end resection. Nat Cell Biol 18, 271–280 (2016). https://doi.org/10.1038/ncb3303

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ncb3303

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