Skip to main content
Log in

Structural and functional characterization of the Spo11 core complex

  • Article
  • Published:

From Nature Structural & Molecular Biology

View current issue Submit your manuscript

Abstract

Spo11, which makes DNA double-strand breaks (DSBs) that are essential for meiotic recombination, has long been recalcitrant to biochemical study. We provide molecular analysis of Saccharomyces cerevisiae Spo11 purified with partners Rec102, Rec104 and Ski8. Rec102 and Rec104 jointly resemble the B subunit of archaeal topoisomerase VI, with Rec104 occupying a position similar to the Top6B GHKL-type ATPase domain. Unexpectedly, the Spo11 complex is monomeric (1:1:1:1 stoichiometry), consistent with dimerization controlling DSB formation. Reconstitution of DNA binding reveals topoisomerase-like preferences for duplex–duplex junctions and bent DNA. Spo11 also binds noncovalently but with high affinity to DNA ends mimicking cleavage products, suggesting a mechanism to cap DSB ends. Mutations that reduce DNA binding in vitro attenuate DSB formation, alter DSB processing and reshape the DSB landscape in vivo. Our data reveal structural and functional similarities between the Spo11 core complex and Topo VI, but also highlight differences reflecting their distinct biological roles.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1: The meiotic DNA DSB core complex.
Fig. 2: Protein–protein interactions within the core complex.
Fig. 3: DNA binding by the core complex analyzed by AFM.
Fig. 4: DNA-binding properties of the core complex.
Fig. 5: Mapping DNA-binding surfaces by hydroxyl radical footprinting.
Fig. 6: Mutations that affect Spo11–DNA binding compromise DSB formation.
Fig. 7: Conformational changes upon DNA binding.
Fig. 8: A mutation that affects Spo11–DNA interaction leads to a redistribution of DSBs.

Similar content being viewed by others

Data availability

Sequence reads and compiled Spo11–oligo and S1–seq maps were deposited at the Gene Expression Omnibus (accession number GSE150315). Source data are provided with this paper.

Code availability

Code for processing Spo11–oligo reads48,49 is available at http://cbio.mskcc.org/public/Thacker_ZMM_feedback/. Code for processing S1–seq reads47 is available at https://github.com/soccin/S1-seq.

References

  1. Bergerat, A. et al. An atypical topoisomerase II from archaea with implications for meiotic recombination. Nature 386, 414–417 (1997).

    Article  CAS  PubMed  Google Scholar 

  2. Keeney, S., Giroux, C. N. & Kleckner, N. Meiosis-specific DNA double-strand breaks are catalyzed by Spo11, a member of a widely conserved protein family. Cell 88, 375–384 (1997).

    Article  CAS  PubMed  Google Scholar 

  3. Corbett, K. D., Benedetti, P. & Berger, J. M. Holoenzyme assembly and ATP-mediated conformational dynamics of topoisomerase VI. Nat. Struct. Mol. Biol. 14, 611–619 (2007).

    Article  CAS  PubMed  Google Scholar 

  4. Graille, M. et al. Crystal structure of an intact type II DNA topoisomerase: insights into DNA transfer mechanisms. Structure 16, 360–370 (2008).

    Article  CAS  PubMed  Google Scholar 

  5. Buhler, C., Lebbink, J. H., Bocs, C., Ladenstein, R. & Forterre, P. DNA topoisomerase VI generates ATP-dependent double-strand breaks with two-nucleotide overhangs. J. Biol. Chem. 276, 37215–37222 (2001).

    Article  CAS  PubMed  Google Scholar 

  6. Lam, I. & Keeney, S. Mechanism and regulation of meiotic recombination initiation. Cold Spring Harb. Perspect. Biol. 7, a016634 (2014).

    Article  PubMed  Google Scholar 

  7. Neale, M. J., Pan, J. & Keeney, S. Endonucleolytic processing of covalent protein-linked DNA double-strand breaks. Nature 436, 1053–1057 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Symington, L. S. End resection at double-strand breaks: mechanism and regulation. Cold Spring Harb. Perspect. Biol. 6, a016436 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Miyoshi, T. et al. A central coupler for recombination initiation linking chromosome architecture to S phase checkpoint. Mol. Cell 47, 722–733 (2012).

    Article  CAS  PubMed  Google Scholar 

  10. Vrielynck, N. et al. A DNA topoisomerase VI-like complex initiates meiotic recombination. Science 351, 939–943 (2016).

    Article  CAS  PubMed  Google Scholar 

  11. Kumar, R., Bourbon, H. M. & de Massy, B. Functional conservation of Mei4 for meiotic DNA double-strand break formation from yeasts to mice. Genes Dev. 24, 1266–1280 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Stanzione, M. et al. Meiotic DNA break formation requires the unsynapsed chromosome axis-binding protein IHO1 (CCDC36) in mice. Nat. Cell Biol. 18, 1208–1220 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Robert, T. et al. The TopoVIB-Like protein family is required for meiotic DNA double strand break formation. Science 351, 943–949 (2016).

    Article  CAS  PubMed  Google Scholar 

  14. Tesse, S. et al. Asy2/Mer2: an evolutionarily conserved mediator of meiotic recombination, pairing, and global chromosome compaction. Genes Dev. 31, 1880–1893 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Keeney, S. Spo11 and the formation of DNA double-strand breaks in meiosis. Genome Dyn. Stab. 2, 81–123 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Malone, R. E. et al. Isolation of mutants defective in early steps of meiotic recombination in the yeast Saccharomyces cerevisiae. Genetics 128, 79–88 (1991).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Kee, K. & Keeney, S. Functional interactions between SPO11 and REC102 during initiation of meiotic recombination in Saccharomyces cerevisiae. Genetics 160, 111–122 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Jiao, K., Salem, L. & Malone, R. Support for a meiotic recombination initiation complex: interactions among Rec102p, Rec104p, and Spo11p. Mol. Cell. Biol. 23, 5928–5938 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Arora, C., Kee, K., Maleki, S. & Keeney, S. Antiviral protein Ski8 is a direct partner of Spo11 in meiotic DNA break formation, independent of its cytoplasmic role in RNA metabolism. Mol. Cell 13, 549–559 (2004).

    Article  CAS  PubMed  Google Scholar 

  20. Kee, K., Protacio, R. U., Arora, C. & Keeney, S. Spatial organization and dynamics of the association of Rec102 and Rec104 with meiotic chromosomes. EMBO J. 23, 1815–1824 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Prieler, S., Penkner, A., Borde, V. & Klein, F. The control of Spo11’s interaction with meiotic recombination hotspots. Genes Dev. 19, 255–269 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Sasanuma, H. et al. Meiotic association between Spo11 regulated by Rec102, Rec104 and Rec114. Nucleic Acids Res. 35, 1119–1133 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Madrona, A. Y. & Wilson, D. K. The structure of Ski8p, a protein regulating mRNA degradation: implications for WD protein structure. Protein Sci. 13, 1557–1565 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Cheng, Z., Liu, Y., Wang, C., Parker, R. & Song, H. Crystal structure of Ski8p, a WD-repeat protein with dual roles in mRNA metabolism and meiotic recombination. Protein Sci. 13, 2673–2684 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Halbach, F., Reichelt, P., Rode, M. & Conti, E. The yeast ski complex: crystal structure and RNA channeling to the exosome complex. Cell 154, 814–826 (2013).

    Article  CAS  PubMed  Google Scholar 

  26. O’Reilly, F. J. & Rappsilber, J. Cross-linking mass spectrometry: methods and applications in structural, molecular and systems biology. Nat. Struct. Mol. Biol. 25, 1000–1008 (2018).

    Article  PubMed  Google Scholar 

  27. Corbett, K. D., Schoeffler, A. J., Thomsen, N. D. & Berger, J. M. The structural basis for substrate specificity in DNA topoisomerase IV. J. Mol. Biol. 351, 545–561 (2005).

    Article  CAS  PubMed  Google Scholar 

  28. Timsit, Y. Local sensing of global DNA topology: from crossover geometry to type II topoisomerase processivity. Nucleic Acids Res. 39, 8665–8676 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Alonso-Sarduy, L., Roduit, C., Dietler, G. & Kasas, S. Human topoisomerase II–DNA interaction study by using atomic force microscopy. FEBS Lett. 585, 3139–3145 (2011).

    Article  CAS  PubMed  Google Scholar 

  30. Wendorff, T. J. & Berger, J. M. Topoisomerase VI senses and exploits both DNA crossings and bends to facilitate strand passage. Elife 7, e31724 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  31. Liu, J., Wu, T. C. & Lichten, M. The location and structure of double-strand DNA breaks induced during yeast meiosis: evidence for a covalently linked DNA-protein intermediate. EMBO J. 14, 4599–4608 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Murakami, H. & Nicolas, A. Locally, meiotic double-strand breaks targeted by Gal4BD-Spo11 occur at discrete sites with a sequence preference. Mol. Cell. Biol. 29, 3500–3516 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Pan, J. et al. A hierarchical combination of factors shapes the genome-wide topography of yeast meiotic recombination initiation. Cell 144, 719–731 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Du, Q., Kotlyar, A. & Vologodskii, A. Kinking the double helix by bending deformation. Nucleic Acids Res. 36, 1120–1128 (2008).

    Article  CAS  PubMed  Google Scholar 

  35. Diaz, R. L., Alcid, A. D., Berger, J. M. & Keeney, S. Identification of residues in yeast Spo11p critical for meiotic DNA double-strand break formation. Mol. Cell. Biol. 22, 1106–1115 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Miller, G. & Hahn, S. A DNA-tethered cleavage probe reveals the path for promoter DNA in the yeast preinitiation complex. Nat. Struct. Mol. Biol. 13, 603–610 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Claeys Bouuaert, C. & Keeney, S. Distinct DNA-binding surfaces in the ATPase and linker domains of MutLγ determine its substrate specificities and exert separable functions in meiotic recombination and mismatch repair. PLoS Genet. 13, e1006722 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  38. Marston, A. L. & Wassmann, K. Multiple duties for spindle assembly checkpoint kinases in meiosis. Front Cell Dev. Biol. 5, 109 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Nichols, M. D., DeAngelis, K., Keck, J. L. & Berger, J. M. Structure and function of an archaeal topoisomerase VI subunit with homology to the meiotic recombination factor Spo11. EMBO J. 18, 6177–6188 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Baudat, F. & Nicolas, A. Clustering of meiotic double-strand breaks on yeast chromosome III. Proc. Natl Acad. Sci. USA 94, 5213–5218 (1997).

    Article  CAS  PubMed  Google Scholar 

  41. Gerton, J. L. et al. Global mapping of meiotic recombination hotspots and coldspots in the yeast Saccharomyces cerevisiae. Proc. Natl Acad. Sci. USA 97, 11383–11390 (2000).

    Article  CAS  PubMed  Google Scholar 

  42. Buhler, C., Borde, V. & Lichten, M. Mapping meiotic single-strand DNA reveals a new landscape of DNA double-strand breaks in Saccharomyces cerevisiae. PLoS Biol. 5, e324 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Blitzblau, H. G., Bell, G. W., Rodriguez, J., Bell, S. P. & Hochwagen, A. Mapping of meiotic single-stranded DNA reveals double-stranded-break hotspots near centromeres and telomeres. Curr. Biol. 17, 2003–2012 (2007).

    Article  CAS  PubMed  Google Scholar 

  44. Wu, T. C. & Lichten, M. Meiosis-induced double-strand break sites determined by yeast chromatin structure. Science 263, 515–518 (1994).

    Article  CAS  PubMed  Google Scholar 

  45. de Massy, B. Initiation of meiotic recombination: how and where? Conservation and specificities among eukaryotes. Annu Rev. Genet 47, 563–599 (2013).

    Article  PubMed  Google Scholar 

  46. Mimitou, E. P., Yamada, S. & Keeney, S. A global view of meiotic double-strand break end resection. Science 355, 40–45 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Mimitou, E. P. & Keeney, S. In Methods in Enzymology Vol. 601 (Eds. Spies, M. & Malkova, A.) 309–330 (Elsevier, 2018).

  48. Thacker, D., Mohibullah, N., Zhu, X. & Keeney, S. Homologue engagement controls meiotic DNA break number and distribution. Nature 510, 241–246 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Mohibullah, N. & Keeney, S. Numerical and spatial patterning of yeast meiotic DNA breaks by Tel1. Genome Res. 27, 278–288 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Bouuaert, C. C. & Keeney, S. Breaking DNA. Science 351, 916–197 (2016).

    Article  CAS  PubMed  Google Scholar 

  51. Yang, J. & Zhang, Y. I-TASSER server: new development for protein structure and function predictions. Nucleic Acids Res. 43, W174–W181 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Bates, A. D., Berger, J. M. & Maxwell, A. The ancestral role of ATP hydrolysis in type II topoisomerases: prevention of DNA double-strand breaks. Nucleic Acids Res. 39, 6327–6339 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Evans, D. H., Li, Y. F., Fox, M. E. & Smith, G. R. A WD repeat protein, Rec14, essential for meiotic recombination in Schizosaccharomyces pombe. Genetics 146, 1253–1264 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Tessé, S., Storlazzi, A., Kleckner, N., Gargano, S. & Zickler, D. Localization and roles of Ski8p in Sordaria macrospora meiosis and delineation of three mechanistically distinct steps of meiotic homolog juxtaposition. Proc. Natl Acad. Sci. USA 100, 12865–12870 (2003).

    Article  PubMed  Google Scholar 

  55. Jolivet, S., Vezon, D., Froger, N. & Mercier, R. Non conservation of the meiotic function of the Ski8/Rec103 homolog in Arabidopsis. Genes Cells 11, 615–622 (2006).

    Article  CAS  PubMed  Google Scholar 

  56. Neale, M. J. & Keeney, S. End-labeling and analysis of Spo11-oligonucleotide complexes in Saccharomyces cerevisiae. Methods Mol. Biol. 557, 183–195 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Johnson, D. et al. Concerted cutting by Spo11 illuminates DNA break mechanisms and initiates gap repair during meiosis. Preprint at bioRxiv https://doi.org/10.1101/2019.12.18.881268 (2019).

  58. Claeys Bouuaert, C., Pu, S., Wang, J., Patel, D. J. & Keeney, S. DNA-dependent macromolecular condensation drives self-assembly of the meiotic DNA break machinery. Preprint at bioRxiv https://doi.org/10.1101/2020.02.21.960245 (2020).

  59. Paiano, J. et al. ATM and PRDM9 regulate SPO11-bound recombination intermediates during meiosis. Nat. Commun. 11, 857 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Yamada, S. et al. Molecular structures and mechanisms of DNA break processing in mouse meiosis. Genes Dev. 34, 806–818 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Kugou, K. et al. Rec8 guides canonical Spo11 distribution along yeast meiotic chromosomes. Mol. Biol. Cell 20, 3064–3076 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Maleki, S., Neale, M. J., Arora, C., Henderson, K. A. & Keeney, S. Interactions between Mei4, Rec114, and other proteins required for meiotic DNA double-strand break formation in Saccharomyces cerevisiae. Chromosoma 116, 471–486 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. McGuffin, L. J., Bryson, K. & Jones, D. T. The PSIPRED protein structure prediction server. Bioinformatics 16, 404–405 (2000).

    Article  CAS  PubMed  Google Scholar 

  64. Richards, F. M. Calculation of molecular volumes and areas for structures of known geometry. Methods Enzymol. 115, 440–464 (1985).

    Article  CAS  PubMed  Google Scholar 

  65. Sebastiaan Winkler, G. et al. Isolation and mass spectrometry of transcription factor complexes. Methods 26, 260–269 (2002).

    Article  CAS  PubMed  Google Scholar 

  66. Erdjument-Bromage, H. et al. Examination of micro-tip reversed-phase liquid chromatographic extraction of peptide pools for mass spectrometric analysis. J. Chromatogr. A. 826, 167–181 (1998).

    Article  CAS  PubMed  Google Scholar 

  67. Yang, B. et al. Identification of cross-linked peptides from complex samples. Nat. Methods 9, 904–906 (2012).

    Article  CAS  PubMed  Google Scholar 

  68. Combe, C. W., Fischer, L. & Rappsilber, J. xiNET: cross-link network maps with residue resolution. Mol. Cell Proteom. 14, 1137–1147 (2015).

    Article  CAS  Google Scholar 

  69. Folta-Stogniew, E. & Williams, K. R. Determination of molecular masses of proteins in solution: implementation of an HPLC size exclusion chromatography and laser light scattering service in a core laboratory. J. Biomol. Tech. 10, 51–63 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Suloway, C. et al. Automated molecular microscopy: the new Leginon system. J. Struct. Biol. 151, 41–60 (2005).

    Article  CAS  PubMed  Google Scholar 

  71. Ludtke, S. J., Baldwin, P. R. & Chiu, W. EMAN: semiautomated software for high-resolution single-particle reconstructions. J. Struct. Biol. 128, 82–97 (1999).

    Article  CAS  PubMed  Google Scholar 

  72. Tang, G. et al. EMAN2: an extensible image processing suite for electron microscopy. J. Struct. Biol. 157, 38–46 (2007).

    Article  CAS  PubMed  Google Scholar 

  73. Scheres, S. H. RELION: implementation of a Bayesian approach to cryo-EM structure determination. J. Struct. Biol. 180, 519–530 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Murakami, H., Borde, V., Nicolas, A. & Keeney, S. Gel electrophoresis assays for analyzing DNA double-strand breaks in Saccharomyces cerevisiae at various spatial resolutions. Methods Mol. Biol. 557, 117–142 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Nicolas, A., Treco, D., Schultes, N. P. & Szostak, J. W. An initiation site for meiotic gene conversion in the yeast Saccharomyces cerevisiae. Nature 338, 35–39 (1989).

    Article  CAS  PubMed  Google Scholar 

  76. Lam, I., Mohibullah, N. & Keeney, S. Sequencing Spo11 oligonucleotides for mapping meiotic DNA double-strand breaks in yeast. Methods Mol. Biol. 1471, 51–98 (2017).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank M. Brendel and the Molecular Cytology core facility at MSKCC for performing the AFM experiments. We thank R. Hendrickson, E. Chang and the Microchemistry and Proteomics core facility at MSKCC for assistance with the XL–MS experiments. We thank K. Liu (S.K. laboratory) for discussions. MSKCC core facilities are supported by National Cancer Institute (NCI) Cancer Center support grant no. P30 CA08748. We thank E. Folta-Stogniew and the Biophysics Resource of Keck Facility at Yale University for the SEC–MALS experiments. The SEC–LS/UV/RI instrumentation was supported by NIH Award Number 1S10RR023748-01. E.P.M. was supported in part by a Helen Hay Whitney Foundation fellowship. Work in the S.K. laboratory was supported principally by the Howard Hughes Medical Institute and in part by NIH grant no. R35 GM118092 (S.K.). Work in the J.M.B. laboratory was funded by NCI grant no. R01-CA0777373 (J.M.B.). C.C.B. was supported in part by funding from the European Research Council under the European Union’s Horizon 2020 research and innovation program (European Research Council grant agreement no. 802525) and from the Fonds National de la Recherche Scientifique (FNRS MIS-Ulysse grant no. F.6002.20) (C.C.B.).

Author information

Authors and Affiliations

Authors

Contributions

C.C.B. and S.K. designed the study and supervised the research. C.C.B. carried out all experiments except those noted below. S.P. generated expression constructs, prepared viruses and purified proteins under the supervision of C.C.B. and S.K. E.P.M. performed S1–seq. S.E.T. performed Spo11–oligo mapping and other phenotypic studies of the F260A mutant and carried out bioinformatics analyses. E.A.-P. performed nsEM experiments with input from J.M.B. J.M.B. generated structural models. C.C.B. and S.K. wrote the paper with input from the other authors. C.C.B., J.M.B. and S.K. secured funding.

Corresponding authors

Correspondence to Corentin Claeys Bouuaert or Scott Keeney.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Peer reviewer reports are available. Beth Moorefield was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

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

Extended data

Extended Data Fig. 1 Co-expression of core complex subunits, size exclusion chromatography and glycerol gradient sedimentation analyses of the core complex.

a, Silver-stained SDS-PAGE gels (top) and anti-Flag western blot (WB; center) of Spo11 complexes after purification on nickel resin. Absence of Rec102, Rec104, or Ski8 leads to poor solubility of Spo11. Bottom: anti-Flag western blot of lysed Sf9 cells showing Spo11 expression levels. Asterisks: C-terminal truncation of Spo11 that retains the affinity tag and interaction with Ski8. b, Size exclusion chromatography of purified core complex with and without MBP tag on Rec102. Silver-stained SDS-PAGE gels of eluted fractions are shown above, with chromatograms from absorption at 280 nm below. c, Glycerol gradient sedimentation of MBP-tagged Spo11 core complexes. The silver-stained SDS-PAGE gel shows fractions collected from the bottom of the gradient. Quantification of protein signal from two independent experiments is shown together with molecular weight markers run on a separate gradient and quantified by Bradford assay. Note: Material in the void volume (panel b) and at the bottom of the glycerol gradient (panel c) lacks Ski8, which is consistent with Ski8 being required for solubility.

Extended Data Fig. 2 2D class averages of nsEM images with different versions of the core complex.

Core complexes without MBP or with MBP fused at the N-terminus of Rec102, Spo11 or Rec104 are shown. A cartoon of the presumed arrangement of the subunits and the position of the MBP signal is shown. With the MBP-tagged Spo11 construct, the signal from MBP is located at a similar position to the Rec102- or Rec104-tagged constructs. This is consistent with the observation that the N-terminus of Spo11 frequently crosslinks with Rec104 (pink lines in Fig. 2a), suggesting that the N-terminus of Spo11, absent from the structural model, is flexible and perhaps directly contacts Rec102/Rec104. The observation of a single MBP signal for all three subunits tested provides further support for the 1:1:1:1 stoichiometry of the core complex. Complexes with MBP-tagged Ski8 were not well behaved and could not be purified.

Extended Data Fig. 3 The interaction between Ski8 and Spo11 is important for the integrity of the complex.

SDS-PAGE analysis of core complexes purified with wild-type Spo11 or the Ski8-interaction deficient Q376A mutant. Equivalent percentages of the total protein purified from similar amounts of Sf9 extract were loaded in each lane, demonstrating the poor yield when the Spo11–Ski8 interaction is compromised.

Extended Data Fig. 4 Intramolecular crosslinks within Ski8 validate the XL-MS results.

a, Ski8 intramolecular crosslinks modeled on the structure of Ski8. The histogram shows the frequency of XL-MS events as a function of distance between the α-carbons (Cα) of the crosslinked lysines (red spheres). The crosslinkable limit of DSS is 27.4 Å. b, Ski8 intramolecular crosslinks modeled on the core complex show that the crosslinked residues are away from the interaction surface with Spo11.

Extended Data Fig. 5 DNA-binding properties of the core complex.

a, Competition experiment using a labeled 25-bp hairpin substrate with 5′-TA overhang in the presence of unlabeled substrates with various overhang configurations. EMSA gel bands of bound labeled substrate are shown. Mean and ranges from two experiments are plotted. The substrate with a 2-nucleotide 5′ overhang is the most effective competitor. b, Competition experiment using a labeled 25-bp hairpin substrate with 5′-TA overhang in the presence of unlabeled competitor substrates with or without 5’ phosphate. Error bars represent ranges from two experiments. c, EMSA of core complex binding to 400-bp mini-circles in the presence or absence of Mg2+. For the top panel, binding reactions contained 5 mM Mg2+ and the gel and electrophoresis buffer contained 0.5 mM Mg2+. For the bottom panel, the binding reactions, gel, and buffer contained 1 mM EDTA.

Extended Data Fig. 6 Affinity purification of different combinations of tagged complexes and comparison of DNA-binding activities.

a, Purification of core complexes that carry combinations of HisFlag (H) and MBP (M) tags on different subunits. All combinations yielded soluble Spo11 (western blot, bottom panel). While the Coomassie-stained gel shown in Fig. 1a suggests that Rec104 may be sub-stoichiometric, the similar relative intensities between MBP-tagged Rec102 and Rec104 in the silver-stained gel (where the MBP tag makes up the majority of each tagged protein’s mass) and anti-MBP western blot indicate that the two subunits have the same stoichiometry (compare lanes 1 with 2, and 6 with 7). The difficulty in purifying soluble Spo11-containing complexes when Rec104 is absent (Extended Data Fig. 1a) further bolsters the inference that the purified core complexes (nearly) always include Rec104. b, Comparison of the DNA-binding activity of core complexes that carried affinity tags on different subunits. All tagged complexes assayed had similar DNA-binding activities.

Extended Data Fig. 7 In vivo analyses of Spo11 DNA-binding mutants.

a, Southern blot analysis of meiotic DSB formation at the CCT6 hotspot in strains expressing wild-type (WT) Spo11 or the K173A or R344A mutant proteins. b, Quantification of DSB formation at the CCT6 hotspot. Error bars represent the range from two experiments. c, Meiotic progression. MI + MII indicates the fraction of cells that have undergone the first or both meiotic divisions, as scored by DAPI staining.

Extended Data Fig. 8 Genome-wide analyses of DSB formation in the F260A mutant.

a, Relative enrichment of the short Spo11-oligo class in F260A. Deproteinized, labeled oligos were separated by denaturing gel electrophoresis. Lane profiles are shown on the right. A 10-nt ladder is plotted in grey. b, c. Reproducibility of DSB maps. Correlations of Spo11-oligo counts (b) and S1-seq counts (c) within hotspots between two biological replicates of Spo11 wild type and F260A are plotted. Pearson’s r between datasets is indicated. d, Changed DSB distribution in F260A is not correlated with hotspot strength. Spo11 hotspots were binned according to oligo counts in wild type. Boxplots show the distribution of Pearson’s r values comparing within-hotspot Spo11-oligo distributions between wild type and F260A, as in Fig. 8e. The thick horizontal bars are medians, box edges are upper and lower quartiles, whiskers indicate values within 1.5 fold of interquartile range, and points are outliers. e, Base composition in S1-seq maps. The big spike in the G map at +2 is partially because this is the complement of the preferred C 5′ of the scissile phosphate, but it is also the first base of the ligation junction (and end-most base after S1 digestion), so the degree to which there is enrichment to the right but not left of the dyad axis probably reflects a modest end-bias in library prep in S1-seq. f, Spo11 preference at the scissile phosphate (dinucleotide indicated by the red circle).

Extended Data Fig. 9 Possible relation of Rec104 to a GHKL fold.

a, Secondary structure predictions for Rec104 and the GHKL domain of Topo VIB were generated by PsiPred63. b, Rec104 model generated by iTasser (green) overlaid on Topo VI. The transducer domain of Topo VI is yellow, the GHKL domain is grey.

Extended Data Fig. 10 Model of Spo11-induced break formation.

a, AFM experiments suggest a model where the core complex binds a DNA duplex, bends it, then traps a second duplex. Perhaps DNA cleavage happens in the context of a trapped DNA junction, similar to Topo VI. After cleavage, Spo11 remains covalently attached to the DNA end through covalent and non-covalent interactions. b, Model of assembly of the DSB machinery. DNA-driven condensation by Rec114–Mei4–Mer2 is proposed to provide a platform that recruits the core complex, where it engages its DNA substrate58. A hypothetical arrangement is shown where each dimer of core complexes captures a pair of DNA duplexes, which, for example, could be sister chromatids. c, Depletion of long oligos in Spo11 mutants with reduced DNA-binding activity is consistent with a model57 where long oligos arise from occlusion of the DNA substrate by multiple Spo11 complexes that reduce access to MRX/Sae2.

Supplementary information

Supplementary Information

Supplementary Tables 2–6.

Reporting Summary

Peer Review Information

Supplementary Table 1

Crosslinking mass spec (XL–MS) data.

Source data

Source Data Fig. 1

Unprocessed gel image

Source Data Fig. 1

Statistical source data for graphs

Source Data Fig. 2

Unprocessed gel images

Source Data Fig. 2

Statistical source data for graphs

Source Data Fig. 3

Statistical source data for graphs

Source Data Fig. 4

Unprocessed gel image

Source Data Fig. 4

Statistical source data for graphs

Source Data Fig. 5

Unprocessed gel images

Source Data Fig. 6

Unprocessed gel image

Source Data Fig. 6

Statistical source data

Source Data Fig. 7

Statistical source data

Source Data Fig. 8

Unprocessed gel images

Source Data Fig. 8

Statistical source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Claeys Bouuaert, C., Tischfield, S.E., Pu, S. et al. Structural and functional characterization of the Spo11 core complex. Nat Struct Mol Biol 28, 92–102 (2021). https://doi.org/10.1038/s41594-020-00534-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-020-00534-w

  • Springer Nature America, Inc.

This article is cited by

Navigation