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:

Protein dynamics during presynaptic-complex assembly on individual single-stranded DNA molecules

Abstract

Homologous recombination is a conserved pathway for repairing double-stranded breaks, which are processed to yield single-stranded DNA overhangs that serve as platforms for presynaptic-complex assembly. Here we use single-molecule imaging to reveal the interplay between Saccharomyces cerevisiae RPA, Rad52 and Rad51 during presynaptic-complex assembly. We show that Rad52 binds RPA–ssDNA and suppresses RPA turnover, highlighting an unanticipated regulatory influence on protein dynamics. Rad51 binding extends the ssDNA, and Rad52–RPA clusters remain interspersed along the presynaptic complex. These clusters promote additional binding of RPA and Rad52. Our work illustrates the spatial and temporal progression of the association of RPA and Rad52 with the presynaptic complex and reveals a new RPA–Rad52–Rad51–ssDNA intermediate, with implications for how the activities of Rad52 and RPA are coordinated with Rad51 during the later stages of recombination.

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: Single-stranded DNA curtain assay for presynaptic-complex assembly.
Figure 2: Individual Rad52 complexes binding to RPA–ssDNA.
Figure 3: Nucleation and growth of Rad52 on RPA–ssDNA.
Figure 4: Rad52 regulates RPA turnover.
Figure 5: Protein dynamics during presynaptic-complex assembly.
Figure 6: Assembly of Rad51–Rad52–RPA–ssDNA presynaptic intermediates.
Figure 7: Model for RPA and Rad52 dynamics during presynaptic-complex assembly.

Similar content being viewed by others

References

  1. Mazon, G., Mimitou, E.P. & Symington, L.S. SnapShot: homologous recombination in DNA double-strand break repair. Cell 142, 646, 646.e1 (2010).

    Article  Google Scholar 

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

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  4. Cromie, G.A., Connelly, J.C. & Leach, D.R. Recombination at double-strand breaks and DNA ends: conserved mechanisms from phage to humans. Mol. Cell 8, 1163–1174 (2001).

    Article  CAS  Google Scholar 

  5. 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  Google Scholar 

  6. Chen, X. et al. The Fun30 nucleosome remodeller promotes resection of DNA double-strand break ends. Nature 489, 576–580 (2012).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  9. 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  Google Scholar 

  10. Wold, M.S. Replication protein A: a heterotrimeric, single-stranded DNA-binding protein required for eukaryotic DNA metabolism. Annu. Rev. Biochem. 66, 61–92 (1997).

    Article  CAS  Google Scholar 

  11. Broderick, S., Rehmet, K., Concannon, C. & Nasheuer, H.P. Eukaryotic single-stranded DNA binding proteins: central factors in genome stability. Subcell. Biochem. 50, 143–163 (2010).

    Article  CAS  Google Scholar 

  12. Mimitou, E.P. & Symington, L.S. DNA end resection:unraveling the tail. DNA Repair (Amst.) 10, 344–348 (2011).

    Article  CAS  Google Scholar 

  13. Lisby, M. & Rothstein, R. Choreography of recombination proteins during the DNA damage response. DNA Repair (Amst.) 8, 1068–1076 (2009).

    Article  CAS  Google Scholar 

  14. 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 

  15. Choi, J.H. et al. Reconstitution of RPA-covered single-stranded DNA-activated ATR-Chk1 signaling. Proc. Natl. Acad. Sci. USA 107, 13660–13665 (2010).

    Article  CAS  Google Scholar 

  16. Maréchal, A. et al. PRP19 Transforms into a sensor of RPA-ssDNA after DNA damage and drives ATR activation via a ubiquitin-mediated circuitry. Mol. Cell 53, 235–246 (2014).

    Article  Google Scholar 

  17. Gasior, S.L., Wong, A.K., Kora, Y., Shinohara, A. & Bishop, D.K. Rad52 associates with RPA and functions with rad55 and rad57 to assemble meiotic recombination complexes. Genes Dev. 12, 2208–2221 (1998).

    Article  CAS  Google Scholar 

  18. Hays, S.L., Firmenich, A.A., Massey, P., Banerjee, R. & Berg, P. Studies of the interaction between Rad52 protein and the yeast single-stranded DNA binding protein RPA. Mol. Cell. Biol. 18, 4400–4406 (1998).

    Article  CAS  Google Scholar 

  19. Oakley, G.G. & Patrick, S.M. Replication protein A: directing traffic at the intersection of replication and repair. Front. Biosci. (Landmark Ed) 15, 883–900 (2010).

    Article  CAS  Google Scholar 

  20. Plate, I. et al. Interaction with RPA is necessary for Rad52 repair center formation and for its mediator activity. J. Biol. Chem. 283, 29077–29085 (2008).

    Article  CAS  Google Scholar 

  21. Conway, A.B. et al. Crystal structure of a Rad51 filament. Nat. Struct. Mol. Biol. 11, 791–796 (2004).

    Article  CAS  Google Scholar 

  22. Ogawa, T., Yu, X., Shinohara, A. & Egelman, E.H. Similarity of the yeast RAD51 filament to the bacterial RecA filament. Science 259, 1896–1899 (1993).

    Article  CAS  Google Scholar 

  23. Heyer, W.D., Ehmsen, K.T. & Liu, J. Regulation of homologous recombination in eukaryotes. Annu. Rev. Genet. 44, 113–139 (2010).

    Article  CAS  Google Scholar 

  24. Symington, L.S. Role of RAD52 epistasis group genes in homologous recombination and double-strand break repair. Microbiol. Mol. Biol. Rev. 66, 630–670 (2002).

    Article  CAS  Google Scholar 

  25. West, S.C. Molecular views of recombination proteins and their control. Nat. Rev. Mol. Cell Biol. 4, 435–445 (2003).

    Article  CAS  Google Scholar 

  26. Bianco, P.R., Tracy, R.B. & Kowalczykowski, S.C. DNA strand exchange proteins: a biochemical and physical comparison. Front. Biosci. 3, D570–D603 (1998).

    Article  CAS  Google Scholar 

  27. Sung, P. Function of yeast Rad52 protein as a mediator between replication protein A and the Rad51 recombinase. J. Biol. Chem. 272, 28194–28197 (1997).

    Article  CAS  Google Scholar 

  28. Shinohara, A. & Ogawa, T. Stimulation by Rad52 of yeast Rad51-mediated recombination. Nature 391, 404–407 (1998).

    Article  CAS  Google Scholar 

  29. New, J.H., Sugiyama, T., Zaitseva, E. & Kowalczykowski, S.C. Rad52 protein stimulates DNA strand exchange by Rad51 and replication protein A. Nature 391, 407–410 (1998).

    Article  CAS  Google Scholar 

  30. Benson, F.E., Baumann, P. & West, S.C. Synergistic actions of Rad51 and Rad52 in recombination and DNA repair. Nature 391, 401–404 (1998).

    Article  CAS  Google Scholar 

  31. Lao, J.P., Oh, S.D., Shinohara, M., Shinohara, A. & Hunter, N. Rad52 promotes postinvasion steps of meiotic double-strand-break repair. Mol. Cell 29, 517–524 (2008).

    Article  CAS  Google Scholar 

  32. McIlwraith, M.J. & West, S.C. DNA repair synthesis facilitates RAD52-mediated second-end capture during DSB repair. Mol. Cell 29, 510–516 (2008).

    Article  CAS  Google Scholar 

  33. Nimonkar, A.V., Sica, R.A. & Kowalczykowski, S.C. Rad52 promotes second-end DNA capture in double-stranded break repair to form complement-stabilized joint molecules. Proc. Natl. Acad. Sci. USA 106, 3077–3082 (2009).

    Article  CAS  Google Scholar 

  34. Sugiyama, T., Kantake, N., Wu, Y. & Kowalczykowski, S.C. Rad52-mediated DNA annealing after Rad51-mediated DNA strand exchange promotes second ssDNA capture. EMBO J. 25, 5539–5548 (2006).

    Article  CAS  Google Scholar 

  35. Lisby, M., Barlow, J.H., Burgess, R.C. & Rothstein, R. Choreography of the DNA damage response: spatiotemporal relationships among checkpoint and repair proteins. Cell 118, 699–713 (2004).

    Article  CAS  Google Scholar 

  36. Lisby, M., Rothstein, R. & Mortensen, U.H. Rad52 forms DNA repair and recombination centers during S phase. Proc. Natl. Acad. Sci. USA 98, 8276–8282 (2001).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  39. Gibb, B. et al. Concentration-dependent exchange of replication protein A on single-stranded DNA revealed by single-molecule imaging. PLoS ONE 9, e87922 (2014).

    Article  Google Scholar 

  40. 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  Google Scholar 

  41. Gibb, B., Silverstein, T.D., Finkelstein, I.J. & Greene, E.C. Single-stranded DNA curtains for real-time single-molecule visualization of protein-nucleic acid interactions. Anal. Chem. 84, 7607–7612 (2012).

    Article  CAS  Google Scholar 

  42. Visnapuu, M.L., Fazio, T., Wind, S. & Greene, E.C. Parallel arrays of geometric nanowells for assembling curtains of DNA with controlled lateral dispersion. Langmuir 24, 11293–11299 (2008).

    Article  CAS  Google Scholar 

  43. Sugiyama, T. & Kowalczykowski, S.C. Rad52 protein associates with replication protein A (RPA)-single-stranded DNA to accelerate Rad51-mediated displacement of RPA and presynaptic complex formation. J. Biol. Chem. 277, 31663–31672 (2002).

    Article  CAS  Google Scholar 

  44. Sugiyama, T. & Kantake, N. Dynamic regulatory interactions of Rad51, Rad52, and replication protein-A in recombination intermediates. J. Mol. Biol. 390, 45–55 (2009).

    Article  CAS  Google Scholar 

  45. Sugiyama, T., New, J.H. & Kowalczykowski, S.C. DNA annealing by RAD52 protein is stimulated by specific interaction with the complex of replication protein A and single-stranded DNA. Proc. Natl. Acad. Sci. USA 95, 6049–6054 (1998).

    Article  CAS  Google Scholar 

  46. 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  Google Scholar 

  47. Singleton, M.R., Wentzell, L.M., Liu, Y., West, S.C. & Wigley, D.B. Structure of the single-strand annealing domain of human RAD52 protein. Proc. Natl. Acad. Sci. USA 99, 13492–13497 (2002).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  49. Graham, J.S., Johnson, R.C. & Marko, J.F. Concentration-dependent exchange accelerates turnover of proteins bound to double-stranded DNA. Nucleic Acids Res. 39, 2249–2259 (2011).

    Article  CAS  Google Scholar 

  50. Sing, C.E., Olvera de la Cruz, M. & Marko, J.F. Multiple-binding-site mechanism explains concentration-dependent unbinding rates of DNA-binding proteins. Nucleic Acids Res. 42, 3783–3791 (2014).

    Article  CAS  Google Scholar 

  51. Sung, P. & Klein, H. Mechanism of homologous recombination: mediators and helicases take on regulatory functions. Nat. Rev. Mol. Cell Biol. 7, 739–750 (2006).

    Article  CAS  Google Scholar 

  52. Wang, X. & Haber, J.E. Role of Saccharomyces single-stranded DNA-binding protein RPA in the strand invasion step of double-strand break repair. PLoS Biol. 2, E21 (2004).

    Article  Google Scholar 

  53. Miyazaki, T., Bressan, D.A., Shinohara, M., Haber, J.E. & Shinohara, A. In vivo assembly and disassembly of Rad51 and Rad52 complexes during double-strand break repair. EMBO J. 23, 939–949 (2004).

    Article  CAS  Google Scholar 

  54. Miné-Hattab, J. & Rothstein, R. Increased chromosome mobility facilitates homology search during recombination. Nat. Cell Biol. 14, 510–517 (2012).

    Article  Google Scholar 

  55. Efron, B. & Tibshirani, R. An Introduction to the Bootstrap (Champman and Hall, New York, 1993).

  56. Antúnez de Mayolo, A. et al. Multiple start codons and phosphorylation result in discrete Rad52 protein species. Nucleic Acids Res. 34, 2587–2597 (2006).

    Article  Google Scholar 

  57. Galletto, R., Amitani, I., Baskin, R.J. & Kowalczykowski, S.C. Direct observation of individual RecA filaments assembling on single DNA molecules. Nature 443, 875–878 (2006).

    Article  CAS  Google Scholar 

  58. Krejci, L. et al. Interaction with Rad51 is indispensable for recombination mediator function of Rad52. J. Biol. Chem. 277, 40132–40141 (2002).

    Article  CAS  Google Scholar 

  59. Busygina, V. et al. Hed1 regulates Rad51-mediated recombination via a novel mechanism. Genes Dev. 22, 786–795 (2008).

    Article  CAS  Google Scholar 

  60. Kwon, Y., Zhao, W. & Sung, P. Biochemical studies on human Rad51-mediated homologous recombination. Methods Mol. Biol. 745, 421–435 (2011).

    Article  CAS  Google Scholar 

  61. Greene, E.C., Wind, S., Fazio, T., Gorman, J. & Visnapuu, M.L. DNA curtains for high-throughput single-molecule optical imaging. Methods Enzymol. 472, 293–315 (2010).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank L. Symington and members of the Greene and Sung laboratories for comments on the manuscript. This research was funded by the US National Institutes of Health grants GM074739 (E.C.G), RO1ES007061 (P.S.) and CA146940 (E.C.G. and P.S.). This work was partially supported by the Nanoscale Science and Engineering Initiative of the US National Science Foundation under award no. CHE-0641523 and by the New York State Office of Science, Technology, and Academic Research (NYSTAR). E.C.G. is supported as an Early Career Scientist by the Howard Hughes Medical Institute.

Author information

Authors and Affiliations

Authors

Contributions

B.G. and L.F.Y. cloned, expressed and purified RPA and Rad52, conducted the single-molecule experiments and analyzed the resulting data. Y.K. and H.N. expressed and purified Rad51, and Y.K. conducted bulk biochemical experiments. E.C.G. supervised the project. B.G. and E.C.G. wrote the manuscript with input from all coauthors.

Corresponding author

Correspondence to Eric C Greene.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Strand-annealing activity assays for fluorescent versions of Rad52.

(a) Schematic of strand annealing assay using complementary 83-nt ssDNA substrates, one of which was radiolabeled (red asterisk). Each substrate was mixed with RPA followed by Rad52, and the two reaction mixes were combined and incubated for the indicated times. (b) Products were resolved by 8% PAGE and (c) quantitated by phosphorimaging (error bars report the standard deviation of 3 experiments).

Supplementary Figure 2 Mediator activity assays for fluorescent versions of Rad52.

(a) Schematic for the mediator assay using a 150-nt ssDNA substrate that was mixed with RPA, Rad52, and Rad52, followed by the addition of a radiolabeled 40-bp dsDNA substrate. (b) Reactions were incubated for an additional 30 minutes, and reaction products were resolved by 8% PAGE and (c) quantitated by phosphorimaging (error bars report the standard deviation of 2-4 experiments). For quantitation, the amount of product formed in the presence of only Rad51 and RPA (lane 3), was subtracted from the amount of product formed in reactions with Rad52.

Supplementary Figure 3 Rad52 does not displace RPA from ssDNA.

A biotinylated 74-nt ssDNA oligo was bound to streptavidin magnetic beads and incubated with saturating amounts of wt RPA (400 nM). Increasing amounts of wt Rad52 were then added to the RPA-bound oligo in the presence of 400 nM free RPA and incubated for 20 minutes at 30 °C. The bound and free fractions were separated by pelleting the beads, and then resolved on an 8-16% SDS-PAGE gel and detected with Coomassie staining. Note that Rfa3 migrates off the gels under these electrophoresis conditions.

Supplementary Figure 4 Rad52 binding to sites occupied by preexisting Rad52 complexes.

(a) Positions of SNAP488-Rad52 (green) bound to a wtRPA-ssDNA complex. (b) Kymograph showing SNAP546-Rad52 (magenta) association with the SNAP488-Rad52/wtRPA/ssDNA complexes. (c) Line graphs showing integrated intensities of SNAP488-Rad52 and average integrated intensity over time for SNAP546-Rad52. The pixel positions for all three panels are co-aligned for direct comparison. (d-f) Same sets of data collected from a different ssDNA molecule.

Supplementary Figure 5 RPA-mCherry bound by wild-type Rad52 resists exchange when chased with wild-type RPA.

(a) Additional examples of kymographs showing RPA-mCherry turnover in the presence of wt RPA. (b) Examples of kymographs showing that the presence of wt Rad52 (0.6 − 1.0 nM) suppresses RPA-mCherry turnover when chased with wt RPA. (c) Examples of integrated RPA-mCherry signal over time (± 0.6 - 1.0 nM Rad52) after chasing with 100 nM wt RPA. The slower decrease in signal observed after the initial phase of rapid turnover is attributed to photo-bleaching.

Supplementary Figure 6 Co-occupancy of RPA, Rad52 and RPA on ssDNA.

Pull down assays performed with (a) ssDNA cellulose and (b) a 74-nt biotinylated oligo bound to streptavidin magnetic beads. The ssDNA was mixed buffer containing wt Rad51 (10 μM), wt Rad52 (1 μM), and wt RPA (0, 25, 50, 100, 200, and 400 nM) for 10 minutes at 37°C. Free and bound fractions were recovered and then resolved on an 8-16% SDS-PAGE gel and detected with Coomassie staining. A beads only negative control (labeled Beads or Beads NC) in (b) shows little protein binding to magnetic beads lacking ssDNA.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–6 (PDF 1615 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gibb, B., Ye, L., Kwon, Y. et al. Protein dynamics during presynaptic-complex assembly on individual single-stranded DNA molecules. Nat Struct Mol Biol 21, 893–900 (2014). https://doi.org/10.1038/nsmb.2886

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nsmb.2886

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