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:

Regulation of the MLH1–MLH3 endonuclease in meiosis

A Publisher Correction to this article was published on 19 January 2021

This article has been updated

Abstract

During prophase of the first meiotic division, cells deliberately break their DNA1. These DNA breaks are repaired by homologous recombination, which facilitates proper chromosome segregation and enables the reciprocal exchange of DNA segments between homologous chromosomes2. A pathway that depends on the MLH1–MLH3 (MutLγ) nuclease has been implicated in the biased processing of meiotic recombination intermediates into crossovers by an unknown mechanism3,4,5,6,7. Here we have biochemically reconstituted key elements of this pro-crossover pathway. We show that human MSH4–MSH5 (MutSγ), which supports crossing over8, binds branched recombination intermediates and associates with MutLγ, stabilizing the ensemble at joint molecule structures and adjacent double-stranded DNA. MutSγ directly stimulates DNA cleavage by the MutLγ endonuclease. MutLγ activity is further stimulated by EXO1, but only when MutSγ is present. Replication factor C (RFC) and the proliferating cell nuclear antigen (PCNA) are additional components of the nuclease ensemble, thereby triggering crossing-over. Saccharomyces cerevisiae strains in which MutLγ cannot interact with PCNA present defects in forming crossovers. Finally, the MutLγ–MutSγ–EXO1–RFC–PCNA nuclease ensemble preferentially cleaves DNA with Holliday junctions, but shows no canonical resolvase activity. Instead, it probably processes meiotic recombination intermediates by nicking double-stranded DNA adjacent to the junction points9. As DNA nicking by MutLγ depends on its co-factors, the asymmetric distribution of MutSγ and RFC–PCNA on meiotic recombination intermediates may drive biased DNA cleavage. This mode of MutLγ nuclease activation might explain crossover-specific processing of Holliday junctions or their precursors in meiotic chromosomes4.

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: MSH4–MSH5 and EXO1(DA) interact with and promote the MLH1–MLH3 endonuclease.
Fig. 2: RFC–PCNA promotes DNA cleavage by the MutLγ–MutSγ–EXO1(DA) ensemble.
Fig. 3: The stimulation of the MLH3 nuclease ensemble by PCNA requires a PIP box motif and is conserved in evolution.

Similar content being viewed by others

Data availability

All relevant data generated or analysed during this study are included in this published article and its Supplementary Information file.

Change history

  • 19 January 2021

    An amendment to this paper has been published and can be accessed via a link at the top of the paper.

References

  1. 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).

    CAS  PubMed  Google Scholar 

  2. Hunter, N. Meiotic recombination: the essence of heredity. Cold Spring Harb. Perspect. Biol. 7, a016618 (2015).

    PubMed  PubMed Central  Google Scholar 

  3. Nishant, K. T., Plys, A. J. & Alani, E. A mutation in the putative MLH3 endonuclease domain confers a defect in both mismatch repair and meiosis in Saccharomyces cerevisiae. Genetics 179, 747–755 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Zakharyevich, K., Tang, S., Ma, Y. & Hunter, N. Delineation of joint molecule resolution pathways in meiosis identifies a crossover-specific resolvase. Cell 149, 334–347 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Zakharyevich, K. et al. Temporally and biochemically distinct activities of Exo1 during meiosis: double-strand break resection and resolution of double Holliday junctions. Mol. Cell 40, 1001–1015 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. De Muyt, A. et al. BLM helicase ortholog Sgs1 is a central regulator of meiotic recombination intermediate metabolism. Mol. Cell 46, 43–53 (2012).

    PubMed  PubMed Central  Google Scholar 

  7. Svetlanov, A., Baudat, F., Cohen, P. E. & de Massy, B. Distinct functions of MLH3 at recombination hot spots in the mouse. Genetics 178, 1937–1945 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Snowden, T., Acharya, S., Butz, C., Berardini, M. & Fishel, R. hMSH4-hMSH5 recognizes Holliday Junctions and forms a meiosis-specific sliding clamp that embraces homologous chromosomes. Mol. Cell 15, 437–451 (2004).

    CAS  PubMed  Google Scholar 

  9. Marsolier-Kergoat, M. C., Khan, M. M., Schott, J., Zhu, X. & Llorente, B. Mechanistic view and genetic control of DNA recombination during meiosis. Mol. Cell 70, 9–20 (2018).

    CAS  PubMed  Google Scholar 

  10. Kadyrov, F. A., Dzantiev, L., Constantin, N. & Modrich, P. Endonucleolytic function of MutLα in human mismatch repair. Cell 126, 297–308 (2006).

    CAS  PubMed  Google Scholar 

  11. Ranjha, L., Anand, R. & Cejka, P. The Saccharomyces cerevisiae Mlh1-Mlh3 heterodimer is an endonuclease that preferentially binds to Holliday junctions. J. Biol. Chem. 289, 5674–5686 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Rogacheva, M. V. et al. Mlh1-Mlh3, a meiotic crossover and DNA mismatch repair factor, is a Msh2-Msh3-stimulated endonuclease. J. Biol. Chem. 289, 5664–5673 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Kadyrova, L. Y., Gujar, V., Burdett, V., Modrich, P. L. & Kadyrov, F. A. Human MutLγ, the MLH1-MLH3 heterodimer, is an endonuclease that promotes DNA expansion. Proc. Natl Acad. Sci. USA 117, 3535–3542 (2020).

    CAS  PubMed  Google Scholar 

  14. Sonntag Brown, M., Lim, E., Chen, C., Nishant, K. T. & Alani, E. Genetic analysis of mlh3 mutations reveals interactions between crossover promoting factors during meiosis in baker’s yeast. G3 3, 9–22 (2013).

    PubMed  Google Scholar 

  15. 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).

    PubMed  PubMed Central  Google Scholar 

  16. Nishant, K. T., Chen, C., Shinohara, M., Shinohara, A. & Alani, E. Genetic analysis of baker’s yeast Msh4-Msh5 reveals a threshold crossover level for meiotic viability. PLoS Genet. 6, e1001083 (2010).

    PubMed  PubMed Central  Google Scholar 

  17. Santucci-Darmanin, S. et al. The DNA mismatch-repair MLH3 protein interacts with MSH4 in meiotic cells, supporting a role for this MutL homolog in mammalian meiotic recombination. Hum. Mol. Genet. 11, 1697–1706 (2002).

    CAS  PubMed  Google Scholar 

  18. Manhart, C. M. et al. The mismatch repair and meiotic recombination endonuclease Mlh1-Mlh3 is activated by polymer formation and can cleave DNA substrates in trans. PLoS Biol. 15, e2001164 (2017).

    PubMed  PubMed Central  Google Scholar 

  19. Kneitz, B. et al. MutS homolog 4 localization to meiotic chromosomes is required for chromosome pairing during meiosis in male and female mice. Genes Dev. 14, 1085–1097 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Flores-Rozas, H. & Kolodner, R. D. The Saccharomyces cerevisiae MLH3 gene functions in MSH3-dependent suppression of frameshift mutations. Proc. Natl Acad. Sci. USA 95, 12404–12409 (1998).

    ADS  CAS  PubMed  Google Scholar 

  21. Lipkin, S. M. et al. MLH3: a DNA mismatch repair gene associated with mammalian microsatellite instability. Nat. Genet. 24, 27–35 (2000).

    CAS  PubMed  Google Scholar 

  22. Wu, Y. et al. A role for MLH3 in hereditary nonpolyposis colorectal cancer. Nat. Genet. 29, 137–138 (2001).

    CAS  PubMed  Google Scholar 

  23. Dherin, C. et al. Characterization of a highly conserved binding site of Mlh1 required for exonuclease I-dependent mismatch repair. Mol. Cell. Biol. 29, 907–918 (2009).

    CAS  PubMed  Google Scholar 

  24. Pluciennik, A. et al. PCNA function in the activation and strand direction of MutLα endonuclease in mismatch repair. Proc. Natl Acad. Sci. USA 107, 16066–16071 (2010).

    ADS  CAS  PubMed  Google Scholar 

  25. Rass, U. et al. Mechanism of Holliday junction resolution by the human GEN1 protein. Genes Dev. 24, 1559–1569 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Bruning, J. B. & Shamoo, Y. Structural and thermodynamic analysis of human PCNA with peptides derived from DNA polymerase-delta p66 subunit and flap endonuclease-1. Structure 12, 2209–2219 (2004).

    CAS  PubMed  Google Scholar 

  27. Lee, S. D. & Alani, E. Analysis of interactions between mismatch repair initiation factors and the replication processivity factor PCNA. J. Mol. Biol. 355, 175–184 (2006).

    CAS  PubMed  Google Scholar 

  28. Liberti, S. E. et al. Bi-directional routing of DNA mismatch repair protein human exonuclease 1 to replication foci and DNA double strand breaks. DNA Repair 10, 73–86 (2011).

    CAS  PubMed  Google Scholar 

  29. Genschel, J. et al. Interaction of proliferating cell nuclear antigen with PMS2 is required for MutLα activation and function in mismatch repair. Proc. Natl Acad. Sci. USA 114, 4930–4935 (2017).

    CAS  PubMed  Google Scholar 

  30. Sanchez, A. A. et al. Mechanism of in vivo activation of the MutLγ-Exo1 complex for meiotic crossover formation. Preprint at https://doi.org/10.1101/2019.12.16.876623 (2019).

    Article  Google Scholar 

  31. El-Shemerly, M., Hess, D., Pyakurel, A. K., Moselhy, S. & Ferrari, S. ATR-dependent pathways control hEXO1 stability in response to stalled forks. Nucleic Acids Res. 36, 511–519 (2008).

    CAS  PubMed  Google Scholar 

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

    ADS  CAS  PubMed  Google Scholar 

  33. Iaccarino, I., Marra, G., Palombo, F. & Jiricny, J. hMSH2 and hMSH6 play distinct roles in mismatch binding and contribute differently to the ATPase activity of hMutSα. EMBO J. 17, 2677–2686 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Anand, R., Pinto, C. & Cejka, P. Methods to study DNA end resection I: recombinant protein purification. Methods Enzymol. 600, 25–66 (2018).

    CAS  PubMed  Google Scholar 

  35. Palombo, F. et al. hMutSβ, a heterodimer of hMSH2 and hMSH3, binds to insertion/deletion loops in DNA. Curr. Biol. 6, 1181–1184 (1996).

    CAS  PubMed  Google Scholar 

  36. Biswas, E. E., Chen, P. H. & Biswas, S. B. Overexpression and rapid purification of biologically active yeast proliferating cell nuclear antigen. Protein Expr. Purif. 6, 763–770 (1995).

    CAS  PubMed  Google Scholar 

  37. Reginato, G., Cannavo, E. & Cejka, P. Physiological protein blocks direct the Mre11-Rad50-Xrs2 and Sae2 nuclease complex to initiate DNA end resection. Genes Dev. 31, 2325–2330 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Arter, M. et al. Regulated crossing-over requires inactivation of Yen1/GEN1 resolvase during meiotic prophase I. Dev. Cell 45, 785–800 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Thacker, D., Lam, I., Knop, M. & Keeney, S. Exploiting spore-autonomous fluorescent protein expression to quantify meiotic chromosome behaviors in Saccharomyces cerevisiae. Genetics 189, 423–439 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Wild, P. et al. Network rewiring of homologous recombination enzymes during mitotic proliferation and meiosis. Mol. Cell 75, 859–874 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Duroc, Y. et al. Concerted action of the MutLβ heterodimer and Mer3 helicase regulates the global extent of meiotic gene conversion. eLife 6, e21900 (2017).

    PubMed  PubMed Central  Google Scholar 

  42. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Chia, M. & van Werven, F. J. Temporal expression of a master regulator drives synchronous sporulation in budding yeast. G3 6, 3553–3560 (2016).

    CAS  PubMed  Google Scholar 

  44. Stahl, F. W. & Lande, R. Estimating interference and linkage map distance from two-factor tetrad data. Genetics 139, 1449–1454 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Matos, J. et al. Dbf4-dependent CDC7 kinase links DNA replication to the segregation of homologous chromosomes in meiosis I. Cell 135, 662–678 (2008).

    CAS  PubMed  Google Scholar 

  46. Borde, V. et al. Histone H3 lysine 4 trimethylation marks meiotic recombination initiation sites. EMBO J. 28, 99–111 (2009).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported by grants from the Swiss National Science Foundation (31003A_17544) and ERC (681-630) to P.C., Institut Curie and CNRS to V.B., Agence Nationale de la Recherche (ANR-15-CE11-0011) to V.B. and J.-B.C., the Novo Nordisk Foundation (NNF15OC0016662) and ERC (724718) to E.R.H., and the Swiss National Science Foundation (155823 and 176108) to J.M. We thank J. Jiricny (ETH Zurich) and members of the Cejka laboratory for helpful comments on the manuscript and N. Hunter for communicating results before publication. We thank ScopeM at ETH Zurich for instrument access.

Author information

Authors and Affiliations

Authors

Contributions

E.C., A.S., R.A. and P.C. planned, performed and analysed the majority of the experiments and wrote the paper. L.R. and A.A. performed most of the experiments with yeast recombinant proteins and electrophoretic mobility shift assays. N.W. performed experiments to define simultaneous DNA binding by MLH1–MLH3 and MSH4–MSH5. J.H. performed experiments with yeast mlh1 and mlh3 variants mutated in PIP-box-like sequences, and analysed the data together with J.M. Chip experiments and Rfc1–Mlh1 and Rfc1–Mlh3 pulldown assays were carried out by C.A., and the data analysed together with V.B. J.-B.C. helped to prepare the MLH1–MLH3 expression construct and designed experiments with the PIP-box peptide. X.A.-G. and E.R.H. prepared the MSH4–MSH5 expression construct. All authors contributed to prepare the final version of the manuscript.

Corresponding author

Correspondence to Petr Cejka.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks Andreas Hochwagen and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Extended data figures and tables

Extended Data Fig. 1 ATP hydrolysis promotes MLH1-MLH3 to nick scDNA.

a, A scheme of MLH1 and MLH3 constructs. The maltose-binding protein (MBP) on MLH3 was cleaved during protein purification. b, Recombinant MLH1-MLH3 used in this study. The 4–15% gradient polyacrylamide gel was stained with Coomassie blue. The experiments in this study used 3 preparations of MLH1-MLH3 that looked similar upon electrophoretic separation and exhibited similar activities. c, Nuclease assay with MLH1-MLH3 and 2.7 kbp-long supercoiled DNA (scDNA) as a substrate. The reaction with 5 mM manganese acetate was incubated without ATP at 37 °C. This experiment was carried out three times with similar results. d, Quantitation of assays such as in c. Averages shown; error bars, SEM; n = 3 independent experiments. e, Nuclease assay with MLH1-MLH3 (300 nM) and 2.7 kbp-long scDNA. Linear DNA was used as a marker. The assay was carried out at 37 °C and contained 5 mM manganese acetate and ATP (0.5 mM). The MLH1-MLH3 nuclease introduces nicks in dsDNA but does not linearize dsDNA. This experiment was carried out three times with similar results. f, Quantitation of nuclease assays with MLH1-MLH3 without or with ATP (0.5 mM), in the presence of manganese (5 mM). Averages shown; error bars, SEM; n = 4 independent experiments. g, Nuclease assay with MLH1-MLH3 and 5 mM magnesium acetate. The reaction buffer contained ATP (0.5 mM). The assay was carried out at 37 °C. The heterodimer exhibits barely detectable nuclease activity in magnesium. This experiment was carried out two times with similar results. h, Nuclease assay with MLH1-MLH3 and various nucleotide cofactors (ADP, ATP and non-hydrolysable ATP analogs ATP-γ-S and AMP-PNP, all 0.5 mM). The assay was carried out at 37 °C with 5 mM manganese acetate. The panel shows a representative experiment. This experiment was carried out four times with similar results. i, Quantitation of nuclease assays such as in panel h, supplemented with various nucleotide co-factors and their analogs (0.5 mM). Averages shown; error bars, SEM; n = 4 independent experiments. j, Purified MLH1-MLH3 variants used in this study. MLH1(EA), MLH1(E34A); MLH3(EA), MLH3(E28A); MLH3(3ND), MLH3(D1223N-Q1224K-E1229K). The 4–15% gradient polyacrylamide gel was stained with Coomassie blue. We used single preparations of the mutant MLH1-MLH3 variants. k, Alignment of MLH1 and MLH3 ATPase motifs. Conserved residues are highlighted in red. Alanine substitutions in MLH3 and MLH1 mutants used in this study are in italics. l, Nuclease assay with wild type MLH1-MLH3 and indicated variants deficient in ATP hydrolysis, without or with ATP (0.5 mM). The assay was carried out at 37 °C, with 5 mM manganese acetate. This experiment was carried out six times with similar results. m, Quantitation of nuclease assays as shown in panel l, without or with ATP (0.5 mM), with either wild type or MLH1-MLH3 variants mutated in conserved ATPase domain residues. Averages shown; error bars, SEM; n = 6 independent experiments. n, Electrophoretic mobility shift assay with indicated MLH1-MLH3 variants, oligonucleotide-based Holliday junction as the substrate, in the absence of ATP and no magnesium (with 3 mM EDTA). Asterisk (*) indicates the position of the 5' radioactive label. A representative experiment is shown at the bottom, a quantitation (averages shown, n = 3 independent experiments; error bars, SEM) at the top. o, Nuclease assays with wild type MLH1-MLH3 on oligonucleotide-based DNA substrates (Holliday junction, HJ and nicked Holliday junction, nicked HJ). The asterisk indicates the position of the 5' radioactive label. The assay was carried out at 37 °C, with 5 mM manganese or magnesium acetate, as indicated, with ATP (1 mM). The products were analysed by 10% native polyacrylamide gel electrophoresis. This experiment was carried out two times with similar results.

Extended Data Fig. 2 Human and yeast MutSγ complexes preferentially bind branched DNA intermediates.

a, Recombinant human MSH4-MSH5 used in this study. We used 3 preparations of MSH4-MSH5 that exhibited similar activities. b, Electrophoretic mobility shift assays with human MSH4-MSH5 and indicated DNA substrates. Asterisk (*) indicates the position of the 5' radioactive label. The assays were carried out in a buffer containing 2 mM magnesium acetate without ATP. The experiments were performed three times. c, Quantitation of DNA binding assays such as shown in panel b. Averages shown; error bars, SEM; n = 3 independent experiments. d, Electrophoretic mobility shift assays with yeast Msh4-Msh5 and indicated DNA substrates. Asterisk (*) indicates the position of the 5' radioactive label. The assays were carried out in a buffer containing 2 mM magnesium acetate without ATP. The experiments were performed two times. e, Quantitation of experiments such as shown in panel d. The lines represent average; n = 2 independent experiments. f, Quantification of electrophoretic mobility shift assays with yeast Msh4-Msh5 and indicated DNA substrates, without magnesium (with 3 mM EDTA). The lines represent average; n = 2 independent experiments.

Extended Data Fig. 3 MutSγ and MutLγ physically interact and moderately stabilize each other at DNA junctions.

a, To investigate the interplay of MutLγ and MutSγ at DNA junctions, we performed electrophoretic mobility shift assays with either or both complexes under more stringent conditions (75 mM NaCl, 2 mM magnesium acetate), separated on 0.6% agarose gels. Under these conditions, MSH4-MSH5 lost the capacity to stably bind Holliday junctions/D-Loops, but could help stabilize the MutSγ-MutLγ complex. The binding of MutLγ alone was not stable, as evidenced by a weak protein-DNA band and the presence of smear in the lanes indicative of complexes that dissociated during electrophoresis. The addition of MutSγ resulted in a moderate stabilization of the protein-DNA complex, and a minor super-shift in electrophoretic mobility of the stable protein-DNA band (indicated by the red and blue arrows). The experiment was performed three times with similar results. b, Electrophoretic mobility shift assays as in panel a, but without magnesium (with 3 mM EDTA). The experiment was performed five times with similar results. c, Quantitation of assays such as shown in panel b. The Y axis indicates relative protein-DNA complex stability, obtained upon dividing the protein-DNA band intensity (see blue or red arrows in panel b) by the intensity of the radioactive signal in the lane above the free substrate band, but below the protein-DNA band. Averages shown; error bars, SEM; n = 5 independent experiments. d, Assays as in a, with human MutLγ and either human or yeast MutSγ. The supershift was observed only when the cognate human complexes were combined. The experiment was performed two times with similar results. e, Electrophoretic mobility shift assays as in a, but with yeast MutLγ and MutSγ complexes. The experiment was performed two times with similar results. f, Protein interaction assays with immobilized MLH1-MLH3 (bait) and MSH4-MSH5 (prey). The 10% polyacrylamide gel was stained with silver. The experiment was performed two times with similar results. g, Protein interaction assays with immobilized human MSH4-MSH5 or yeast Msh4-Msh5 that were used as baits, and human MLH1-MLH3 (prey). The eluted proteins were analysed by silver staining. Although interaction between yeast Msh4-Msh5 and human MLH1-MLH3 was still detected, it was weaker than the interaction between the cognate MSH4-MSH5 and MLH1-MLH3 complexes. The experiment was performed two times with similar results. h, Protein interaction assay with immobilized yeast Msh4-Msh5 (bait) and yeast Mlh1-Mlh3 (prey). The eluted proteins were analysed by western blotting. The experiment was performed two times with similar results. i, Electrophoretic mobility shift assays with MLH1-MLH3 and MSH4-MSH5, as indicated, and oligonucleotide-based Holliday junction DNA substrate. 32P-labelled λDNA/HindIII digest was used as a marker. The DNA-bound MLH1-MLH3 and MSH4-MSH5 species migrate high up on the agarose gel where the resolution capacity is limited. The experiment was performed two times with similar results. j, Electrophoretic mobility shift assay with yeast Ku70-Ku80 heterodimer and Holliday junction DNA substrate. Ku bound the dsDNA ends of the four Holliday junction arms, resulting in up to 4 heterodimers bound to the DNA substrate (lanes 5-7). Comparison with λ DNA/HindIII and panel i revealed that the Ku-DNA complex migrates much faster than DNA-bound MLH1-MLH3 and MSH4-MSH5. This suggests that multiple units of MLH1-MLH3 and MSH4-MSH5 bind DNA. The experiment was performed two times with similar results.

Extended Data Fig. 4 MSH4-MSH5 promotes DNA cleavage by MLH1-MLH3, but the complex does not exhibit resolvase activity.

a, Quantitation of kinetic nuclease assays with MLH1-MLH3 (50 nM) without or with MSH4-MSH5 (50 nM) using 5.6 kbp-long scDNA. The assays were carried out at 30 °C in the presence of 5 mM manganese acetate and 2 mM ATP. Averages shown; error bars, SEM; n = 3 independent experiments. b, Nuclease assays with MSH4-MSH5 and either wild type MLH1-MLH3 or nuclease-dead MLH1-MLH3 (D1223N-Q1224K-E1229K, 3ND). The assays were carried out at 30 °C in the presence of 5 mM manganese acetate and 0.5 mM ATP. The experiment was performed three times with similar results. c, Quantitation of nuclease assays with various MLH1-MLH3 and MSH4-MSH5 concentrations, as indicated. The assays were carried out at 30 °C in the presence of 5 mM manganese acetate and 0.5 mM ATP. Averages shown; error bars, SEM, n = 3 independent experiments. The efficiency of nuclease cleavage was generally dependent on the concentrations used. When using 50 nM MLH1-MLH3, the maximal cleavage efficiency was achieved together with 50 nM MSH4-MSH5, no further increase when using 100 nM MSH4-MSH5 was observed. This suggests that both heterodimers may form a stoichiometric complex. Vice versa, when using 50 nM MSH4-MSH5, a further increase of DNA cleavage was observed when MLH1-MLH3 concentrations exceeded 50 nM, which is in agreement with the capacity of MLH1-MLH3 to cleave DNA on its own. d, Quantitation of nuclease assays with MLH1-MLH3 and MSH4-MSH5, as indicated, in the presence of various nucleotide co-factors or their analogs (2 mM). The assays were carried out at 30 °C in the presence of 5 mM manganese acetate. Averages shown; error bars, SEM; n = 6 independent experiments. e, Representative nuclease assays with MSH4-MSH5 and variants of MLH1-MLH3 deficient in ATP hydrolysis, as indicated. The assays were carried out at 30 °C in the presence of 5 mM manganese acetate and 0.5 mM ATP. The experiment was performed three times with similar results. f, Representative nuclease assays with MLH1-MLH3 and variants of MSH4-MSH5 deficient in ATP hydrolysis, as indicated. The assays were carried out at 30 °C in the presence of 5 mM manganese acetate and 0.5 mM ATP. The experiment was performed three times with similar results. g, Recombinant MSH4-MSH5 and its variants used in this study. MSH4(G685A), MSH4(GA); MSH5(G597A), MSH5(GA). The 4–15% gradient polyacrylamide gel was stained with Coomassie blue. We used a single preparation of the MSH4-MSH5 mutant variants. h, Quantitation of electrophoretic mobility shift assays with MSH4-MSH5 and its ATPase motif mutant variants. Oligonucleotide-based Holliday junction was used as the substrate. Asterisk (*) indicates the position of the 5' radioactive label. ATP was not included in the binding buffer. The mutations did not affect the capacity of MSH4-MSH5 to bind DNA. Averages shown; error bars, SEM; n = 3 independent experiments. i, Nuclease reactions were carried out with yeast or human MutSγ and MutLγ complexes, as indicated (50 nM), with 2.7 kbp-long scDNA substrate. While human MutSγ promoted DNA cleavage by human MutLγ (compare lanes 2 and 3), yeast MutSγ did not notably promote DNA cleavage by human MutLγ (compare lanes 2 and 5), and reciprocally, human MutSγ did not promote DNA cleavage by yeast MutLγ (compare lanes 7 and 8). The experiment was performed two times with similar results. j, Quantitation of nuclease assays with human and yeast MutSγ and MutLγ complexes as in panel i, but with 10.3 kbp-long scDNA substrate. Averages shown; error bars, SEM; n = 3 independent experiments. k, Cleavage of pIRbke8mut cruciform DNA (inverted repeats folding back to form a Holliday junction structure) by MutSγ and MutLγ complexes. The quantitation below the lanes represents an average from two independent experiments. Simultaneous cleavage of both strands at the junction point would lead to linear DNA. No linear DNA was observed with MutSγ and MutLγ, indicating a lack of canonical resolvase activity. The experiment was performed nine times with similar results. l, Representative nuclease assays with indicated proteins and oligonucleotide-based Holliday junction DNA. Asterisk (*) indicates the position of the 5' radioactive label. No DNA cleavage was observed, indicating a lack of structure-specific DNA cleavage activity on the oligonucleotide-based substrate. The products were analysed by 15% denaturing polyacrylamide gel electrophoresis. The experiment was performed two times with similar results.

Extended Data Fig. 5 MutSβ but not MutSα stimulates MutLγ to a similar extent as MutSγ.

a, Coomassie-stained polyacrylamide gel showing recombinant MutSβ (MSH2-MSH3). We used one MSH2-MSH3 preparation in this study. b, Coomassie-stained polyacrylamide gel showing recombinant MutSα (MSH2-MSH6) used in this study. We used one MSH2-MSH6 preparation in this study. c, Nuclease assays with MLH1-MLH3, MSH4-MSH5, and MSH2-MSH3 or MSH2-MSH6, as indicated. The assays were carried out at 30 °C in the presence of 5 mM manganese acetate and 0.5 mM ATP. A representative experiment is shown at the bottom, a quantitation (averages shown; n = 3 independent experiments; error bars, SEM) at the top.

Extended Data Fig. 6 Stimulation of the nuclease activity of MutSγ-MutLγ by EXO1(D173A).

a, Recombinant EXO1(D173A), used in this study. The 4–15% gradient polyacrylamide gel was stained with Coomassie blue. We used three EXO1(D173A) preparations in this work. b, Nuclease assays with MLH1-MLH3 and MSH4-MSH5, as indicated, without (left) or with EXO1(D173A) (right). The assays were carried out at 30 °C in the presence of 5 mM manganese acetate and 0.5 mM ATP. A representative experiment is shown at the bottom, a quantitation (averages shown; n = 3 independent experiments; error bars, SEM) at the top. c, Nuclease assays with MLH1-MLH3 and/or EXO1(D173A), as indicated. The assays were carried out at 30 °C in the presence of 5 mM manganese acetate and 0.5 mM ATP. A representative experiment is shown at the bottom, a quantitation (averages shown; n = 4 independent experiments; error bars, SEM) at the top. EXO1(DA) does not promote the nuclease of MLH1-MLH3 alone. The limited DNA cleavage in lane 3 likely results from residual nuclease activity of EXO1(D173A) that becomes apparent at high protein concentrations (100 nM) in the presence of manganese. d, Quantitation of electrophoretic mobility shift assays with MLH1-MLH3, MSH4-MSH5 and EXO1(D173A), as indicated. The protein-DNA species were resolved in 1% agarose gels. Averages shown; error bars, SEM; n = 5 independent experiments. EXO1(D173A) did not notably affect DNA binding of MLH1-MLH3 and MSH4-MSH5. e, Nuclease assays with MLH1-MLH3, MSH4-MSH5 with either human EXO1(D173A) or yeast Exo1(D173A), as indicated. The assays were carried out at 30 °C in the presence of 5 mM manganese acetate and 0.5 mM ATP. A representative experiment is shown at the bottom, a quantitation (averages shown; n = 5 independent experiments; error bars, SEM) at the top. f, Nuclease assays with MLH1-MLH3, MSH2-MSH3 and EXO1(D173A), as indicated. The assays were carried out at 30 °C in the presence of 5 mM manganese acetate and 0.5 mM ATP. A representative experiment is shown at the bottom, a quantitation (averages shown; n = 3 independent experiments; error bars, SEM) at the top.

Extended Data Fig. 7 RFC-PCNA promote the nuclease activity of the MutSγ-MutLγ-EXO1(DA) ensemble.

a, Recombinant human and yeast RFC and PCNA used in this study. The 4–15% gradient polyacrylamide gel was stained with Coomassie blue. We used two yeast RFC and PCNA preparations, and one human RFC preparation in this work. b, Nuclease assays with scDNA and indicated proteins (all 50 nM, except human PCNA, 100 nM) were carried out with 5 mM magnesium acetate and 2 mM ATP at 37 °C. A representative experiment is shown at the bottom, a quantitation (averages shown; n = 5 independent experiments; error bars, SEM) at the top. c, Experiments as in panel b, comparing the efficacy of human and yeast RFC as a part of the MLH3 nuclease ensemble. Averages shown; n = 4 independent experiments; error bars, SEM. d, Nuclease reactions containing MLH1-MLH3 (50 nM), MSH4-MSH5 (50 nM), EXO1(D173A) (50 nM) and yRFC-hPCNA (50-100 nM, respectively) (column 1), without MSH4-MSH5 (column 2) or without EXO1(D173A) (column 3). Reactions were carried out with 5 mM magnesium acetate and 2 mM ATP at 37 °C. Averages shown; error bars, SEM; n = 5 independent experiments. e, Kinetic nuclease assays with MLH1-MLH3 (50 nM), MSH4-MSH5 (50 nM), EXO1(D173A) (50 nM) and yRFC-hPCNA (50-100 nM, respectively), as indicated. Reactions were carried out with 5 mM magnesium acetate and 2 mM ATP at 37 °C. Averages shown; error bars, SEM; n = 5 independent experiments. f, Nuclease assays with MLH1-MLH3 (50 nM), MSH4-MSH5 (50 nM), EXO1(D173A) (50 nM) and hRFC-hPCNA (50-100 nM, respectively), as indicated, with supercoiled (left) or relaxed DNA (right). Reactions were carried out with 5 mM magnesium acetate and 2 mM ATP at 37 °C. Shown is a representative experiment. RFC-PCNA do not stimulate the cleavage of relaxed DNA. The experiment was performed three times with similar results. g, Nuclease assays with MLH1-MLH3, MSH4-hMSH5, EXO1(D173A) without or with yRFC-hPCNA, as indicated. The assays were carried out at 37 °C in the presence of 5 mM manganese acetate and 2 mM ATP. A representative experiment is shown at the bottom, a quantitation (averages shown; n = 3 independent experiments; error bars, SEM) at the top. Without magnesium, no stimulation of DNA cleavage by RFC-PCNA was observed. h, Nuclease reactions with MLH1-MLH3 (50 nM), MSH4-MSH5 (50 nM), EXO1(D173A) (50 nM) and yRFC-hPCNA (50-100 nM, respectively), as indicated. Reactions were carried out with 5 mM magnesium acetate and 2 mM ATP at 37 °C. Averages shown; error bars, SEM; n = 5 independent experiments. i, Nuclease assays with MLH1-MLH3 (50 nM), MSH4-MSH5 (50 nM), EXO1(D173A) (50 nM) and yRFC-hPCNA (50-100 nM, respectively) and 5 mM magnesium acetate, either with no nucleotide co-factor (lane 2), with ATP (2 mM, lane 3) or ADP (2 mM, lane 4). ATP is strictly required for DNA cleavage by the nuclease ensemble. The experiment was performed four times with similar results. j, Representative nuclease assays with MLH1-MLH3 (50 nM), MSH4-MSH5 (50 nM), EXO1(D173A) (50 nM) and yRFC-hPCNA (50-100 nM, respectively), lane 2. Lanes 3-7 contain instead MLH1-MLH3 or MSH4-MSH5 variants deficient in ATP hydrolysis, as indicated. See Fig. 1d, e for the specific mutations. Reactions were carried out with 5 mM magnesium acetate and 2 mM ATP at 37 °C. The experiment was performed four times with similar results. k, Nuclease assays with indicated oligonucleotide-based substrates carried out at 37 °C in the presence of 5 mM magnesium acetate and 2 mM ATP. All proteins 30 nM, as indicated. Asterisk (*) indicates the position of the 5' radioactive label. The reaction products were analysed on a 15% denaturing polyacrylamide gel. No DNA cleavage was observed. The experiment was performed two times with similar results.

Extended Data Fig. 8 PIP box-like motifs in EXO1, MLH3 and MLH1 facilitate the stimulatory effect of RFC-PCNA on the hMLH3 nuclease ensemble.

a, The MLH1P-MLH3P variant (see Fig. 3b) is not impaired in Holliday junction-binding. Electrophoretic mobility shift assay was carried out with 5 ng/reaction dsDNA competitor and 3 mM EDTA (no magnesium). Asterisk (*) indicates the position of the 5' radioactive label. The experiment was carried out three times with similar results. b, The MLH1P and MLH3P variant combinations are not impaired in nuclease activity without or with MSH4-MSH5 and EXO1(D173A) in the absence of RFC-PCNA. The nuclease assays were performed with 5 mM manganese acetate and 2 mM ATP at 37 °C. Averages shown; error bars, SEM, n = 3 independent experiments. c, Nuclease assays with MSH4-MSH5 (50 nM), EXO1(D173A) (50 nM) and yRFC-hPCNA (50-100 nM), and a respective MLH1-MLH3 variant, as indicated (see Fig. 3b). Mutations in the PIP-box like motif reduce the stimulation of the nuclease ensemble by RFC-PCNA. The assays were carried out with 5 mM magnesium acetate and 2 mM ATP at 37 °C. Averages shown; error bars, SEM, n = 5 independent experiments. d, The EXO1P(D173A) variant with mutated PIP-box motif (see Fig. 3b) is not affected in its ability to promote the nuclease of MLH1-MLH3 and MSH4-MSH5 (without RFC-PCNA). The assays were carried out with 5 mM manganese acetate and 2 mM ATP at 37 °C. Averages shown; error bars, SEM, n = 4 independent experiments. e, The EXO1P(D173A) variant with mutated PIP-box motif (see Fig. 3b), in complex with MLH1-MLH3 and MSH4-MSH5 impairs the stimulatory function of yRFC-hPCNA (50-100 nM). The assays were carried out with 5 mM magnesium acetate and 2 mM ATP at 37 °C. Averages shown; error bars, SEM, n = 5 independent experiments.

Extended Data Fig. 9 RFC-PCNA promote meiotic recombination in yeast cells.

a, Spore viability upon tetrad microdissection, analysed in the wild type strain, mlh1Δ and mlh3Δ, and in strains complemented with a construct expressing untagged Mlh1P (Q572A-L575A-F578A) or Mlh3P (Q293A-V296A-F300A) at the endogenous chromosomal locus. At least 156 spores from 2 independent experiments were analysed for each genotype. b, western blot analysis of Mlh1P expression in yeast. TCA extracts were prepared from exponentially proliferating SK1 strains expressing MLH1, MLH1-FLAG or MLH1P-FLAG from the endogenous gene locus. The PIP-box-like mutation affects the stability of the FLAG-tagged Mlh1 protein. Blots were probed with anti-FLAG antibody (Sigma, F7425). Crm1 is a protein normalization control. Asterisk denotes a cross-reacting band. The western blot was carried out three times from two different TCA extractions with similar results. c, western blot analysis of Mlh3P expression in yeast. As in b, but with MLH3, MLH3-FLAG or MLH3P-FLAG constructs. Blots were probed with anti-FLAG antibodies: Sigma F7425 (left panel); A8592 (right panel). Crm1 is a protein normalization control. Mlh1-FLAG and Mlh3P-FLAG showed comparable expression levels. Asterisks denote cross-reacting bands. The western blot was carried out three (left panel) or two (right panel) times from two different TCA extractions with similar results. d, A pulldown of TAP-tagged yeast Rfc1-5 and associated proteins from meiotic cell extracts from pCUP1-IME1 cells 5 h 30 min after the induction of meiosis. The presence of Mlh1-HA and Mlh3-Myc in the TEV eluate was analysed by western blotting. This experiment was carried out once. e, Rfc1-TAP levels at the three indicated meiotic DSB hotspots relative to a negative control site (NFT1) were assessed by ChIP and qPCR in ndt80∆-arrested cells after 7 h in meiosis. Mlh3 is not required for the recruitment of RFC to the meiotic DSB hotspots. MLH3: VBD2136; mlh3∆: VBD2137. Averages shown; n = 2 independent experiments.

Extended Data Fig. 10 A possible model for biased resolution of recombination intermediates by the MLH3 nuclease ensemble.

Meiotic dsDNA breaks (a) are resected (1) and the resulting DNA overhang invades matching DNA on a homologous chromosome (2). The unstable D-Loop intermediates (b) are stabilized by MSH4-MSH5 (3), DNA synthesis by RFC-PCNA-Polδ (4) and branch migration (5), leading to more stable structures termed single-end invasions (c). This is followed by a second end capture (6), and more DNA synthesis (7) leading to precursors of double Holliday junctions (d) and later matured double Holliday junctions (e). As a result of the previous steps, MSH4-MSH5 and RFC-PCNA may be present asymmetrically at the (d) or (e) intermediates at the junctions points or their vicinity. The asymmetric presence of the co-factors then directs and stimulates the biased DNA cleavage (9) of (d) or (e) structures by MLH1-MLH3-EXO1. Upon final processing (10) and ligation (11), the ultimate result is a DNA crossover characterized by reciprocal exchange of the DNA arms of the recombining chromosomes.

Supplementary information

Supplementary Table

Supplementary Table 1. Oligonucleotides used in this study.

Reporting Summary

Supplementary Table

Supplementary Table 2. Saccharomyces cerevisiae strains used in this study. All strains are SK1 derivatives.

Supplementary Data

Gel source data (uncropped gel images).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cannavo, E., Sanchez, A., Anand, R. et al. Regulation of the MLH1–MLH3 endonuclease in meiosis. Nature 586, 618–622 (2020). https://doi.org/10.1038/s41586-020-2592-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-020-2592-2

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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