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:

POLθ-mediated end joining is restricted by RAD52 and BRCA2 until the onset of mitosis

A Publisher Correction to this article was published on 27 October 2021

This article has been updated

Abstract

BRCA2-mutant cells are defective in homologous recombination, making them vulnerable to the inactivation of other pathways for the repair of DNA double-strand breaks (DSBs). This concept can be clinically exploited but is currently limited due to insufficient knowledge about how DSBs are repaired in the absence of BRCA2. We show that DNA polymerase θ (POLθ)-mediated end joining (TMEJ) repairs DSBs arising during the S phase in BRCA2-deficient cells only after the onset of the ensuing mitosis. This process is regulated by RAD52, whose loss causes the premature usage of TMEJ and the formation of chromosomal fusions. Purified RAD52 and BRCA2 proteins both block the DNA polymerase function of POLθ, suggesting a mechanism explaining their synthetic lethal relationships. We propose that the delay of TMEJ until mitosis ensures the conversion of originally one-ended DSBs into two-ended DSBs. Mitotic chromatin condensation might further serve to juxtapose correct break ends and limit chromosomal fusions.

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: Formation and repair of spontaneous DSBs throughout the cell cycle.
Fig. 2: TMEJ in HR-deficient cells is delayed until mitosis.
Fig. 3: RAD52 locates to resected DSBs to prevent their repair in G2-phase BRCA2 mutants.
Fig. 4: RAD52 prevents the premature usage of TMEJ in BRCA2 mutants.
Fig. 5: RAD52 and BRCA2 both suppress TMEJ.
Fig. 6: RAD52 and BRCA2 inhibit the polymerase function of POLθ.
Fig. 7: RAD52 and BRCA2 inhibit POLθ DNA synthesis by distinct mechanisms.

Similar content being viewed by others

Data availability

Source data are provided with this study. All data supporting the findings of this study are available from the corresponding author on reasonable request. Source data are provided with this paper.

Change history

References

  1. Roy, R., Chun, J. & Powell, S. N. BRCA1 and BRCA2: different roles in a common pathway of genome protection. Nat. Rev. Cancer 12, 68–78 (2012).

    Article  CAS  Google Scholar 

  2. Prakash, R., Zhang, Y., Feng, W. & Jasin, M. Homologous recombination and human health: the roles of BRCA1, BRCA2, and associated proteins. Cold Spring Harb. Perspect. Biol. 7, a01660 (2015).

    Article  Google Scholar 

  3. Higgins, G. S. & Boulton, S. J. Beyond PARP–POLθ as an anticancer target. Science 359, 1217–1218 (2018).

    Article  CAS  PubMed  Google Scholar 

  4. Ashworth, A. & Lord, C. J. Synthetic lethal therapies for cancer: what’s next after PARP inhibitors? Nat. Rev. Clin. Oncol. 15, 564–576 (2018).

    Article  CAS  PubMed  Google Scholar 

  5. Jensen, R. B., Carreira, A. & Kowalczykowski, S. C. Purified human BRCA2 stimulates RAD51-mediated recombination. Nature 467, 678–683 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Liu, J., Doty, T., Gibson, B. & Heyer, W. D. Human BRCA2 protein promotes RAD51 filament formation on RPA-covered single-stranded DNA. Nat. Struct. Mol. Biol. 17, 1260–1262 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Ceccaldi, R. et al. Homologous-recombination-deficient tumours are dependent on Polθ-mediated repair. Nature 518, 258–262 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Mateos-Gomez, P. A. et al. Mammalian polymerase θ promotes alternative NHEJ and suppresses recombination. Nature 518, 254–257 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Wood, R. D. & Doublié, S. DNA polymerase θ (POLQ), double-strand break repair, and cancer. DNA Repair 44, 22–32 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Wyatt, D. W. et al. Essential roles for polymerase θ-mediated end joining in the repair of chromosome breaks. Mol. Cell 63, 662–673 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Feng, Z. et al. Rad52 inactivation is synthetically lethal with BRCA2 deficiency. Proc. Natl Acad. Sci. USA 108, 686–691 (2011).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Mazina, O. M., Keskin, H., Hanamshet, K., Storici, F. & Mazin, A. V. Rad52 inverse strand exchange drives RNA-templated DNA double-strand break repair. Mol. Cell 67, 19–29 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  15. Spies, J. et al. 53BP1 nuclear bodies enforce replication timing at under-replicated DNA to limit heritable DNA damage. Nat. Cell Biol. 21, 487–497 (2019).

    Article  CAS  PubMed  Google Scholar 

  16. Hengel, S. R. et al. Small-molecule inhibitors identify the RAD52-ssDNA interaction as critical for recovery from replication stress and for survival of BRCA2 deficient cells. eLife 5, e14740 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  17. Huang, F. et al. Targeting BRCA1-and BRCA2-deficient cells with RAD52 small molecule inhibitors. Nucleic Acids Res. 44, 4189–4199 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Hanamshet, K., Mazina, O. M. & Mazin, A. V. Reappearance from obscurity: mammalian Rad52 in homologous recombination. Genes 7, 63 (2016).

    Article  PubMed Central  Google Scholar 

  19. Löbrich, M. et al. H2AX foci analysis for monitoring DNA double-strand break repair: strengths, limitations and optimization. Cell Cycle 9, 662–669 (2010).

    Article  PubMed  Google Scholar 

  20. Pommier, Y., Sun, Y., Huang, S. Y. N. & Nitiss, J. L. Roles of eukaryotic topoisomerases in transcription, replication and genomic stability. Nat. Rev. Mol. Cell Biol. 17, 703–721 (2016).

    Article  CAS  PubMed  Google Scholar 

  21. Juhász, S., Elbakry, A., Mathes, A. & Löbrich, M. ATRX promotes DNA repair synthesis and sister chromatid exchange during homologous recombination. Mol. Cell 71, 11–24 (2018).

    Article  PubMed  Google Scholar 

  22. Spies, J. et al. Nek1 regulates Rad54 to orchestrate homologous recombination and replication fork stability. Mol. Cell 62, 903–917 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Howlett, N. G. et al. Biallelic inactivation of BRCA2 in Fanconi anemia. Science 297, 606–609 (2002).

    Article  CAS  PubMed  Google Scholar 

  24. Bennardo, N., Cheng, A., Huang, N. & Stark, J. M. Alternative-NHEJ is a mechanistically distinct pathway of mammalian chromosome break repair. PLoS Genet. 4, e1000110 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  25. Han, J. et al. BRCA2 antagonizes classical and alternative nonhomologous end-joining to prevent gross genomic instability. Nat. Commun. 8, 1470 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Kelso, A. A., Lopezcolorado, F. W., Bhargava, R. & Stark, J. M. Distinct roles of RAD52 and POLQ in chromosomal break repair and replication stress response. PLoS Genet. 15, e1008319 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Stasiak, A. Z. et al. The human Rad52 protein exists as a heptameric ring. Curr. Biol. 10, 337–340 (2000).

    Article  CAS  PubMed  Google Scholar 

  28. Ducy, M. et al. The tumor suppressor PALB2: inside out. Trends Biochem. Sci. 44, 226–240 (2019).

    Article  CAS  PubMed  Google Scholar 

  29. Mateos-Gomez, P. A. et al. The helicase domain of Polθ counteracts RPA to promote alt-NHEJ. Nat. Struct. Mol. Biol. 24, 1116–1123 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Grimme, J. M. et al. Human Rad52 binds and wraps single-stranded DNA and mediates annealing via two hRad52-ssDNA complexes. Nucleic Acids Res. 38, 2917–2930 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Shinohara, A., Shinohara, M., Ohta, T., Matsuda, S. & Ogawa, T. Rad52 forms ring structures and co-operates with RPA in single-strand DNA annealing. Genes Cells 3, 145–156 (1998).

    Article  CAS  PubMed  Google Scholar 

  32. Von Nicolai, C., Ehlén, Å., Martin, C., Zhang, X. & Carreira, A. A second DNA binding site in human BRCA2 promotes homologous recombination. Nat. Commun. 7, 12813 (2016).

    Article  Google Scholar 

  33. Yang, H. et al. BRCA2 function in DNA binding and recombination from a BRCA2–DSS1–ssDNA structure. Science 297, 1837–1848 (2002).

    Article  CAS  PubMed  Google Scholar 

  34. Zámborszky, J. et al. Loss of BRCA1 or BRCA2 markedly increases the rate of base substitution mutagenesis and has distinct effects on genomic deletions. Oncogene 36, 746–755 (2017).

    Article  PubMed  Google Scholar 

  35. Deng, S. K., Gibb, B., De Almeida, M. J., Greene, E. C. & Symington, L. S. RPA antagonizes microhomology-mediated repair of DNA double-strand breaks. Nat. Struct. Mol. Biol. 21, 405–412 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Oshima, J., Pae, C., Huang, S., Campisi, J. & Schiestl, R. H. Lack of WRN results in extensive deletion at nonhomologous joining ends. Cancer Res. 62, 547–551 (2002).

    CAS  PubMed  Google Scholar 

  37. Ahnesorg, P., Smith, P. & Jackson, S. P. XLF interacts with the XRCC4–DNA ligase IV complex to promote DNA nonhomologous end-joining. Cell 124, 301–313 (2006).

    Article  CAS  PubMed  Google Scholar 

  38. Kegel, P., Riballo, E., Kühne, M., Jeggo, P. A. & Löbrich, M. X-irradiation of cells on glass slides has a dose doubling impact. DNA Repair 6, 1692–1697 (2007).

    Article  CAS  PubMed  Google Scholar 

  39. Zahn, K., Jensen, R. B., Wood, R. D. & Doublié, S. Human DNA polymerase θ harbors a DNA end-trimming activity critical for DNA repair. Mol. Cell 81, 1534–1547 (2021).

  40. Seki, M., Marini, F. & Wood, R. D. POLQ (Pol θ), a DNA polymerase and DNA-dependent ATPase in human cells. Nucleic Acids Res. 31, 6117–6126 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Jensen, R. Purification of Recombinant 2XMBP tagged human proteins from human cells. Methods Mol. Biol. 1176, 209–217 (2014).

    Article  CAS  PubMed  Google Scholar 

  42. Rothenberg, E., Grimme, J. M., Spies, M. & Ha, T. Human Rad52-mediated homology search and annealing occurs by continuous interactions between overlapping nucleoprotein complexes. Proc. Natl Acad. Sci. USA 105, 20274–20279 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Zhang, X. P. & Heyer, W. D. Quality control of purified proteins involved in homologous recombination. Methods Mol. Biol. 745, 329–343 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Le, H. P. et al. DSS1 and ssDNA regulate oligomerization of BRCA2. Nucleic Acids Res. 48, 7818–7833 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Carreira, A. et al. The BRC repeats of BRCA2 modulate the DNA-binding selectivity of RAD51. Cell 136, 1032–1043 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Sneeden, J. L., Grossi, S. M., Tappin, I., Hurwitz, J. & Heyer, W. D. Reconstitution of recombination-associated DNA synthesis with human proteins. Nucleic Acids Res. 41, 4913–4925 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Binz, S. K., Dickson, A. M., Haring, S. J. & Wold, M. S. Functional assays for replication protein A (RPA). Methods Enzymol. 409, 11–38 (2006).

    Article  CAS  PubMed  Google Scholar 

  48. Van Komen, S., Macris, M., Sehorn, M. G. & Sung, P. Purification and assays of Saccharomyces cerevisiae homologous recombination proteins. Methods Enzymol. 408, 445–463 (2006).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We thank J. Stark for providing the U2OS EJ2 reporter and POLθ/RAD52 KO cell lines, and A. Löwer and P. Jeggo for helpful discussions. We thank R. Weimer, C. Braun and M. P. Lowery for technical assistance, and D. Deckbar and A. Taubmann for experimental and intellectual contributions during the early stages of the project. Work in the M.L. laboratory is supported by the Deutsche Forschungsgemeinschaft (DFG, project ID 393547839—SFB 1361) and the Bundesministerium für Bildung und Forschung (grant nos 02NUK054C, 02NUK050B and 02NUK042D), work in the W.D.H. laboratory is supported by the National Institutes of Health (grant nos GM58015 and GM137751), and work in the R.D.W. laboratory is supported by the National Institutes of Health (grant nos CA247773 and CA193124) and the J. Ralph Meadows Chair in Carcinogenesis Research. This research was also supported by the National Institutes of Health award CA187561 to J.L. and by core services supported by P30 CA93373.

Author information

Authors and Affiliations

Authors

Contributions

M.L.A., M.E., H.P.L., A.G., J.L. and A.C.G. performed experiments and interpreted data. S.B. and R.D.W. provided reagents. M.L. and W.D.H. conceived the experiments and wrote the paper aided by M.L.A., M.E., H.P.L., A.G., J.L., A.C.G., S.B. and R.D.W.

Corresponding author

Correspondence to Markus Löbrich.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Cell Biology thanks the anonymous reviewers for their contribution to the peer review of this work. Peer reviewer reports are available.

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

Extended data

Extended Data Fig. 1 Repair of spontaneous breaks by TMEJ in HR-deficient cells is delayed until mitosis.

a, Spontaneous γH2AX foci in G1 and G2 HeLa cells. Cells were transfected with siRNAs, labelled with EdU for 1 h and foci were immediately analysed after labelling in EdU cells (microscopic dot blot on the left indicates the evaluated cell populations) (n = 4 independent experiments, except siRAD51 + siPOLθ: n = 3 independent experiments). b, Spontaneous γH2AX foci in HeLa cells treated with a second POLθ siRNA (siPOLθ_6). (i) Cells were labelled as in Fig. 1a and EdUBrdU+ cells were analysed (n = 3 independent experiments, except siCTRL/siBRCA2 at 12 h: n = 4 independent experiments). (ii) Cells were labelled and analysed as in (a) (n = 3 independent experiments, except siCTRL/siBRCA2/siRAD51: n = 4 independent experiments). The data without siPOLθ_6 in (i) are also part of Fig. 2a,b whilst the data without siPOLθ_6 in (ii) are also shown in panel (a) of the present figure. c, Spontaneous γH2AX foci in fibroblasts. The same analysis as in (b) was performed with 82-6 hTERT (WT) cells (for (i): n = 4 independent experiments, for (ii): n = 3 independent experiments). d, Spontaneous chromatid breaks in G2 and mitotic HeLa cells. Cells were transfected with siRNAs and analysed by premature chromosome condensation (PCC) for G2 cells or a 2-h colcemid treatment for mitotic cells. Chromatid breaks were quantified per spread and normalized to 70 chromosomes (n = 3 independent experiments). e, Lagging chromosomes in ana/telophase HeLa cells. Cells were transfected with siRNAs and ana/telophases were identified by their morphology seen in the DAPI and phospho-histone H3(S10) (pH3) stainings (n = 4 independent experiments). f, Micronuclei in G1 HeLa cells. Cells were transfected with siRNAs, labelled as in Fig. 1a and EdUBrdU+ cells were analysed in G1 (n = 4 independent experiments). All data show the mean ± s.e.m. Individual experiments, each derived from 40 (S/G2 foci, G2 spreads and lagging chromosomes), 60 (M spreads), 100 cells (G1 foci) or 200 cells (micronuclei), are shown as dots. *: P < 0.05, **: P < 0.01; ***: P < 0.001, ns: non-significant (One-way ANOVA). The exact P values are provided as source data. Source data are available online.

Source data

Extended Data Fig. 2 Repair of exogenous breaks by TMEJ in HR-deficient cells is delayed until mitosis.

a, γH2AX foci after CPT treatment in S/G2 and G1 HeLa cells. Cells were transfected with siRNAs and treated with EdU and 20 nM CPT for 1 h. EdU+ cells were in S during the CPT treatment and progressed to G2 and G1 after 12 h. Foci were analysed in these cells at the indicated times (microscopic dot blot on the left indicates the evaluated cell populations) (n = 4 independent experiments). b, γH2AX foci after IR in G2 HeLa cells. Cells were transfected with siRNAs, labelled with EdU for 3 h and irradiated with 2 Gy X-rays, after which BrdU was added. EdU+BrdU cells were in G2 at the time of IR and partially progressed to G1 after 12 h, at which time they were analysed (microscopic dot blot on the left indicates the evaluated cell populations) (n = 3 independent experiments). c, γH2AX foci in HeLa cells treated with PARPi for 24 h. Cells were transfected with siRNAs, treated with a PARP inhibitor (Olaparib, 1 µM) for 24 h, pulse-labelled with EdU for the last hour and γH2AX foci were then analysed in EdU G2, mitotic or G1 cells (n = 3 independent experiments). All data show the mean ± s.e.m. Individual experiments, each derived from 40 (S/G2), 80 (Ana/Telo) or 100 cells (G1), are shown as dots. **: P < 0.01; ***: P < 0.001, ns: non-significant (One-way ANOVA). The exact P values are provided as source data. Source data are available online.

Source data

Extended Data Fig. 3 RAD52 locates to resected DSBs to prevent their repair in G2-phase BRCA2 mutants.

a, Diagram of the GFP–RAD52 protein: GFP tag (green), RAD52 self-binding domain (orange), RPA-binding domain (purple), RAD51-binding domain (yellow) and nuclear localization signal (NLS, red). Expression levels of endogenous vs. tagged RAD52 were assessed by immunoblotting. GAPDH was used as loading control. The example immunofluorescence (IF) image shows the variability in GFP–RAD52 expression between different cells (scale bar, 10 µm). For foci analysis, cells with very high or very low GFP–RAD52 expression were excluded. b, Kinetics of γH2AX, RAD51 and GFP–RAD52 foci in G2 HeLa cells stably expressing GFP–RAD52. Cells were labelled with EdU, irradiated with 2 Gy X-rays and EdU G2 cells were analysed (n = 5 independent experiments, except 0.25 h and 5 h: n = 3 or n = 2 independent experiments, respectively). c, Co-localization studies of RAD52, RAD51, pRPA and γH2AX in G2 HeLa GFP–RAD52 cells. Cells were labelled with EdU, irradiated with 2 Gy X-rays and EdU G2 cells were analysed (n = 2 independent experiments, except co-localization of RAD52/RAD51: n = 3 independent experiments). Example IF images are shown in the right panel (scale bar, 5 µm). d, Spontaneous γH2AX foci in S/G2 HeLa cells. Cells were transfected with siRNAs, labelled as in Fig. 1a and EdUBrdU+ cells were analysed (n = 4 independent experiments). e, γH2AX foci after IR in G2 HeLa cells. Cells were transfected with siRNAs, labelled with EdU, irradiated with 2 Gy X-rays and EdU G2 cells were analysed (n = 3 independent experiments). f, γH2AX foci after IR in G2 fibroblast cells. Experiment was performed as in (e), but using 82-6 hTERT (WT) cells (n = 3 independent experiments). Data without siBRCA2 are the same as in Fig. 3e. All data show the mean ± s.e.m, except for (b) and the conditions in (c) with n = 2 independent experiments for which only the mean is shown. Individual experiments, each derived from 40 cells (except (c): 20 cells), are shown as dots. *: P < 0.05; ***: P < 0.001, ns: non-significant (One-way ANOVA). The exact P values are provided as source data. Source data are available online.

Source data

Extended Data Fig. 4 RAD52 limits the premature usage of TMEJ in BRCA2 mutants to prevent the formation of chromosomal fusions.

a, γH2AX foci after IR in G2 HeLa cells. Cells were transfected with siRNAs, labelled with EdU, irradiated with 2 Gy X-rays and EdU G2 cells were analysed (n = 4 independent experiments). b, γH2AX foci after IR in G1 and G2 HeLa cells. Cells were transfected with siRNAs, labelled with EdU, irradiated with 2 Gy X-rays and EdU G1 and G2 cells were analysed (n = 2 independent experiments for G1 and n = 3 independent experiments for G2, except siLIG4/siLIG4 + siRAD52/siBRCA2 + siLIG4: n = 2 independent experiments). The G2 data without siLIG4 are also shown in Extended Data Fig. 3e. c, γH2AX foci after IR in G2 fibroblasts lacking XLF. 2BN hTERT (XLF deficient: XLF*) cells were transfected with siRNAs, labelled with EdU, irradiated with 2 Gy X-rays and EdU G2 cells were analysed (n = 3 independent experiments). XLF mutants have much higher levels of unrepaired breaks compared with WT fibroblasts due to their defect in canonical non-homologous end joining, a major pathway for repairing IR-induced DSBs in G2 (compare to Extended Data Fig. 3f). d, Chromatid breaks and fusions after IR and CPT in G2 HeLa cells. Cells were transfected with siRNAs and treated with IR (2 Gy X-rays) or CPT (20 nM, 1 h). Chromosome spreads were obtained by PCC 8-10 h post IR/CPT treatment. Chromatid breaks and fusions were quantified per spread and normalized to 70 chromosomes (n = 4 independent experiments, except siBRCA2 + siPOLθ/siBRCA2 + siRAD52 + siPOLθ: n = 3 independent experiments). Examples of chromosome spreads with breaks (red arrows) and fusions (blue arrows) in irradiated HeLa cells are shown (scale bar, 5 µm). All data show the mean ± s.e.m., except for conditions in (b) with n = 2 independent experiments for which only the mean is shown. Individual experiments, each derived from 30 (G2 spreads) or 40 cells (G1 and G2), are shown as dots. *: P < 0.05; **: P < 0.01; ***: P < 0.001, ns: non-significant (One-way ANOVA). The exact P values are provided as source data. Source data are available online.

Source data

Extended Data Fig. 5 Elevated end joining frequencies arise after co-depletion of BRCA2 and RAD52.

a, Diagram of the reporter assay monitoring end joining (EJ) events utilizing micro-homologies (modified from ref. 24). The U2OS EJ2 reporter cells carry a GFP cassette which is preceded by an I-SceI recognition sequence and stop codons in all reading frames flanked by two 8 bp regions of homology (shown in red). Repair events that result in the removal of the stop codons and maintain the GFP reading frame will result in expression of functional GFP. The majority of such events use the 8 bp homologous regions for annealing, leading to a deletion of 35 bp and the formation of a XcmI restriction site24. b, Frequency of GFP + U2OS EJ2 reporter cells. Cells were transfected with siRNAs and treated with triamcinolone acetonide (TA) (200 nM) to induce I-SceI localization to the nucleus. (i) Upon break induction, the percentage of GFP + cells was quantified using microscopic analysis and normalized to the siCTRL sample (n = 3 independent experiments). In addition, GFP − and GFP + cells were sorted, followed by DNA extraction and PCR amplification across the break site. The final products were then subjected to a restriction test as follows: (ii) Molecular analysis of control GFP − - and GFP + -sorted cells. For GFP − cells, the PCR products across the break site can be efficiently cut by I-SceI but not by XcmI. For GFP + cells, the products can be cut by XcmI but not by I-SceI, indicating that the original recognition sequence is lost and a XcmI restriction site is formed due to micro-homology usage at the 8 bp repeat. (iii) Molecular analysis of GFP + -sorted cells. For all siRNA conditions, the PCR products across the break site can be efficiently cut by XcmI but not by I-SceI, as seen in the GFP + control in (ii). All data show the mean ± s.e.m. Individual experiments, each derived from 10,000 events, are shown as dots. **: P < 0.01; ***: P < 0.001, ns: non-significant (One-way ANOVA). The exact P values are provided as source data. Source data are available online.

Source data

Extended Data Fig. 6 Absent or deregulated TMEJ causes distinct mitotic aberrations.

a, Spontaneous γH2AX foci in G1 and G2 U2OS cells. U2OS WT and KO cells were transfected with siRNAs, labelled with EdU for 1 h and foci were immediately analysed after labelling in EdU cells in G1 and G2 (n = 3 independent experiments). b, Alternative representation of the data in Fig. 4e. U2OS WT and KO cells were transfected with siRNAs and ana/telophase cells were analysed for the presence of lagging chromosomes and anaphase bridges. Cells were identified by their morphology seen in the DAPI and phospho-histone H3(S10) (pH3) stainings. The percentage of cells showing 0 to ≥4 spontaneous lagging chromosomes or 0 to ≥3 anaphase bridges is represented. (combined raw data from n = 3 independent experiments). c, Spontaneous anaphase bridges in mitotic HeLa cells. Cells were transfected with siRNAs and anaphase cells were analysed for the presence of anaphase bridges. Cells were identified by their morphology seen in the DAPI and phospho-histone H3(S10) (pH3) stainings (n = 3 independent experiments). All data show the mean ± s.e.m. Individual experiments, each derived from 40 (G2 foci and anaphase bridges), 80 (lagging chromosomes) or 100 cells (G1), are shown as dots. ***: P < 0.001, ns: non-significant (One-way ANOVA). The exact P values are provided as source data. Source data are available online.

Source data

Extended Data Fig. 7 The RPA and SELF, but not the RAD51, binding domains of RAD52 are essential for suppressing TMEJ.

a, Diagram of GFP–RAD52 WT and mutant proteins (left panel). Tables show amino acid positions of the deleted RAD52 domains and the molecular weights of the resulting GFP–RAD52 mutant proteins (right panel). b, (i) Expression levels of GFP–RAD52 mutant proteins were analysed by immunoblotting in two independent clones for each mutant (#1 and 2). (ii) Depletion of endogenous RAD52 with the single siRNA was confirmed by immunoblotting in the control cell line (GFP only). Vinculin was used as loading control. c, γH2AX and pRPA foci in G2 HeLa cells carrying different GFP–RAD52 constructs (clone #2 of each mutant). Cells were transfected with siRNAs, labelled with EdU, irradiated with 2 Gy X-rays and EdUGFP + G2 cells were analysed (n = 3 independent experiments). All data show the mean ± s.e.m. Individual experiments, each derived from 40 cells, are shown as dots. *: P < 0.05; **: P < 0.01; ***: P < 0.001, ns: non-significant (One-way ANOVA). The exact P values are provided as source data. Source data are available online.

Source data

Extended Data Fig. 8 RAD52 and BRCA2 inhibit the polymerase function of POLθ via their ssDNA binding modes.

a, Gel images of purified human BRCA2, RAD52, RAD51, RPA and POLθ. 200 ng BRCA2, 2 µg RAD52, 1 µg RAD51, 2 µg RPA and 500 ng POLθ were loaded. The visible contaminants for POLθ include HSP70, HSP90 and tubulin. The gels shown are single experiments. b, Titration of POLθ and E. coli PolI in the primer-extension assay using the 80 nucleotide 5’-overhang substrate. c, Titration of human RAD52 in the primer-extension assay using the 80 nucleotide 5’-overhang substrate. d, Titration of BRCA2 in the primer-extension assay using the 80 nucleotide 5’-overhang substrate. e, Effect of human RAD52 (hRAD52) and S. cerevisiae Rad52 (yRad52) on POLθ-mediated DNA synthesis in the presence or absence of 4 nM human RPA and 100 µM ATP. A titration of hRAD52 in the primer-extension assay using the 80 nucleotide 5’-overhang substrate was performed. A single concentration for yRad52 was also tested. f, Effect of BRCA2 on POLθ-mediated DNA synthesis in the presence or absence of 4 nM human RPA (hRPA) or S. cerevisiae RPA (yRPA) and 100 µM ATP. A titration of BRCA2 in the primer-extension assay using the 80 nucleotide 5’-overhang substrate was performed. g, Titration of hRAD52 and yRad52 in the primer-extension assay using the 80 nucleotide 5’-overhang. The schemes of the reactions are shown in the diagrams at the top. Protein concentrations are indicated. DNAs were analysed on 10% non-denaturing gels. Extension values are normalized as described in the Methods section. For (e,f): All the percentages of extended products are normalized to POLθ-only reaction (±ATP). Representative gels are shown for the quantitation in Fig. 6b, which reports mean ± s.e.m. (n = 3 independent experiments). For (b): Result from a single experiment is shown. The result was consistent with results from two additional experiments performed under slightly varying conditions. For (c,d,g): Results from single experiments are shown that were performed once under the condition shown, and once with slightly varying conditions with consistent results. Source data are available online.

Source data

Extended Data Fig. 9 RPA and RAD51 don’t inhibit the polymerase function of POLθ.

a, Titration of ssDNA binding proteins (human RPA: hRPA, S. cerevisiae RPA: yRPA and the phage ssDNA binding protein: T4 gp32) in the primer-extension assay using the 80 nucleotide 5’-overhang substrate. The ratio between protein and DNA is indicated as molar ratio. Human or yeast RPA fully covers the available ssDNA at the 2:1 ratio, and gp32 fully covers all ssDNA at the 20:1 ratio. b, Titration of human RAD51 in the primer-extension assay using the 80 nucleotide 5’-overhang substrate with the 5’-Cy5-labelled 20-mer. RAD51 fully covers the available ssDNA at the 1 RAD51 per 3 nucleotide ratio (1:3). The schemes of the reactions are shown in the diagrams at the top. Protein concentrations are indicated. DNAs were analysed on 10% non-denaturing gels for (a) and on 15% denaturing gels for (b). Extension values and unutilized primer values are normalized as described in the Methods section. Results from single experiments are shown. The result in (b) was consistent with the result from one additional experiment performed under slightly varying conditions. Source data are available online.

Source data

Extended Data Fig. 10 Immunoblotting and qPCR analyses confirming siRNA efficiencies.

a, Depletion of RAD52 and BRCA2 in HeLa cells, as well as in 82-6 hTERT (WT) and HSC-62 hTERT (BRCA2 mutant: BRCA2*) fibroblasts, was confirmed by immunoblotting. GAPDH, KU70 and KU80 were used as loading controls (data shown represent 3 independent experiments for HeLa and WT, and 2 independent experiments for BRCA2*). b, Depletion of POLθ was confirmed by qPCR, due to the lack of reliable antibodies for immunoblotting. The standard siRNA used for most experiments (siPOLθ_1; referred to as siPOLθ within the manuscript), as well as an independent siRNA (siPOLθ_6) were tested (n = 3 independent experiments, except siCTRL/siPOLθ_1 in fibroblasts: n = 4 independent experiments). All data show the mean ± s.e.m. Individual experiments are shown as dots. **: P < 0.01; ***: P < 0.001 (One-way ANOVA was used to compare data from siPOLθ_1, while an unpaired two-tailed t-test was used for comparing samples from siPOLθ_6). The exact P values are provided as source data. c,d, Depletion of RAD51, RAD52, BRCA2 and LIG4 in HeLa cells was confirmed by immunoblotting. Vinculin, GAPDH and KU80 were used as loading controls. e, Depletion of BRCA2, RAD54, RAD52 and RAD51 (left), as well as PALB2 (right) in HeLa GFP–RAD52 cells was confirmed by immunoblotting. KU80 and GAPDH were used as loading controls. Source data are available online. Data shown represent 2 independent experiments in c and e, and 3 independent experiments in d.

Source data

Supplementary information

Reporting Summary

Supplementary Table 1

List of oligonucleotides used in this study.

Peer Review Information

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 6

Uncropped gel images.

Source Data Fig. 7

Statistical source data.

Source Data Fig. 7

Uncropped gel images.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 3

Uncropped gel images.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 5

Uncropped gel images.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 7

Uncropped gel images.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 8

Uncropped gel images.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 9

Uncropped gel images.

Source Data Extended Data Fig. 10

Statistical source data.

Source Data Extended Data Fig. 10

Uncropped gel images.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Llorens-Agost, M., Ensminger, M., Le, H.P. et al. POLθ-mediated end joining is restricted by RAD52 and BRCA2 until the onset of mitosis. Nat Cell Biol 23, 1095–1104 (2021). https://doi.org/10.1038/s41556-021-00764-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-021-00764-0

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