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.

  • Letter
  • Published:

Mechanism for remodelling of the cell cycle checkpoint protein MAD2 by the ATPase TRIP13

Abstract

The maintenance of genome stability during mitosis is coordinated by the spindle assembly checkpoint (SAC) through its effector the mitotic checkpoint complex (MCC), an inhibitor of the anaphase-promoting complex (APC/C, also known as the cyclosome)1,2. Unattached kinetochores control MCC assembly by catalysing a change in the topology of the β-sheet of MAD2 (an MCC subunit), thereby generating the active closed MAD2 (C-MAD2) conformer3,4,5. Disassembly of free MCC, which is required for SAC inactivation and chromosome segregation, is an ATP-dependent process driven by the AAA+ ATPase TRIP13. In combination with p31comet, an SAC antagonist6, TRIP13 remodels C-MAD2 into inactive open MAD2 (O-MAD2)7,8,9,10. Here, we present a mechanism that explains how TRIP13–p31comet disassembles the MCC. Cryo-electron microscopy structures of the TRIP13–p31comet–C-MAD2–CDC20 complex reveal that p31comet recruits C-MAD2 to a defined site on the TRIP13 hexameric ring, positioning the N terminus of C-MAD2 (MAD2NT) to insert into the axial pore of TRIP13 and distorting the TRIP13 ring to initiate remodelling. Molecular modelling suggests that by gripping MAD2NT within its axial pore, TRIP13 couples sequential ATP-driven translocation of its hexameric ring along MAD2NT to push upwards on, and simultaneously rotate, the globular domains of the p31comet–C-MAD2 complex. This unwinds a region of the αA helix of C-MAD2 that is required to stabilize the C-MAD2 β-sheet, thus destabilizing C-MAD2 in favour of O-MAD2 and dissociating MAD2 from p31comet. Our study provides insights into how specific substrates are recruited to AAA+ ATPases through adaptor proteins and suggests a model of how translocation through the axial pore of AAA+ ATPases is coupled to protein remodelling.

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: Overall structures of the apo and TRIP13–p31–substrate complexes.
Fig. 2: Interaction between p31comet and TRIP13.
Fig. 3: Interaction between C-MAD2NT and the TRIP13 pore loops.
Fig. 4: Sequential catalytic cycles of TRIP13 remodel MAD2.
Fig. 5: Differences between basal and activated states of TRIP13–p31–substrate.

Similar content being viewed by others

References

  1. Musacchio, A. The molecular biology of spindle assembly checkpoint signaling dynamics. Curr. Biol. 25, R1002–R1018 (2015).

    Article  PubMed  CAS  Google Scholar 

  2. .Alfieri, C., Zhang, S. & Barford, D. Visualizing the complex functions and mechanisms of the anaphase promoting complex/cyclosome (APC/C). Open Biol. (2017).

  3. Kulukian, A., Han, J. S. & Cleveland, D. W. Unattached kinetochores catalyze production of an anaphase inhibitor that requires a Mad2 template to prime Cdc20 for BubR1 binding. Dev. Cell 16, 105–117 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  4. Faesen, A. C. et al. Basis of catalytic assembly of the mitotic checkpoint complex. Nature 542, 498–502 (2017).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  5. Ji, Z., Gao, H., Jia, L., Li, B. & Yu, H. A sequential multi-target Mps1 phosphorylation cascade promotes spindle checkpoint signaling. eLife 6, e22513 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  6. Habu, T., Kim, S. H., Weinstein, J. & Matsumoto, T. Identification of a MAD2-binding protein, CMT2, and its role in mitosis. EMBO J. 21, 6419–6428 (2002).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  7. Eytan, E. et al. Disassembly of mitotic checkpoint complexes by the joint action of the AAA-ATPase TRIP13 and p31comet. Proc. Natl Acad. Sci. USA 111, 12019–12024 (2014).

    Article  ADS  PubMed  CAS  Google Scholar 

  8. Wang, K. et al. Thyroid hormone receptor interacting protein 13 (TRIP13) AAA-ATPase is a novel mitotic checkpoint-silencing protein. J. Biol. Chem. 289, 23928–23937 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  9. Teichner, A. et al. p31comet promotes disassembly of the mitotic checkpoint complex in an ATP-dependent process. Proc. Natl Acad. Sci. USA 108, 3187–3192 (2011).

    Article  ADS  PubMed  Google Scholar 

  10. Westhorpe, F. G., Tighe, A., Lara-Gonzalez, P. & Taylor, S. S. p31comet-mediated extraction of Mad2 from the MCC promotes efficient mitotic exit. J. Cell Sci. 124, 3905–3916 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  11. Reddy, S. K., Rape, M., Margansky, W. A. & Kirschner, M. W. Ubiquitination by the anaphase-promoting complex drives spindle checkpoint inactivation. Nature 446, 921–925 (2007).

    Article  ADS  PubMed  CAS  Google Scholar 

  12. Foster, S. A. & Morgan, D. O. The APC/C subunit Mnd2/Apc15 promotes Cdc20 autoubiquitination and spindle assembly checkpoint inactivation. Mol. Cell 47, 921–932 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  13. Mansfeld, J., Collin, P., Collins, M. O., Choudhary, J. S. & Pines, J. APC15 drives the turnover of MCC-CDC20 to make the spindle assembly checkpoint responsive to kinetochore attachment. Nat. Cell Biol. 13, 1234–1243 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  14. Uzunova, K. et al. APC15 mediates CDC20 autoubiquitylation by APC/C(MCC) and disassembly of the mitotic checkpoint complex. Nat. Struct. Mol. Biol. 19, 1116–1123 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  15. Jia, L. et al. Defining pathways of spindle checkpoint silencing: functional redundancy between Cdc20 ubiquitination and p31comet. Mol. Biol. Cell 22, 4227–4235 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  16. Eytan, E., Sitry-Shevah, D., Teichner, A. & Hershko, A. Roles of different pools of the mitotic checkpoint complex and the mechanisms of their disassembly. Proc. Natl Acad. Sci. USA 110, 10568–10573 (2013).

    Article  ADS  PubMed  Google Scholar 

  17. Ye, Q. et al. TRIP13 is a protein-remodeling AAA+ ATPase that catalyzes MAD2 conformation switching. eLife 4, e07367 (2015).

    Article  PubMed Central  CAS  Google Scholar 

  18. Ye, Q. et al. The AAA+ ATPase TRIP13 remodels HORMA domains through N-terminal engagement and unfolding. EMBO J. 36, 2419–2434 (2017).

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  19. Yang, M. et al. p31comet blocks Mad2 activation through structural mimicry. Cell 131, 744–755 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  20. Ma, H. T. & Poon, R. Y. C. TRIP13 regulates both the activation and inactivation of the spindle-assembly checkpoint. Cell Reports 14, 1086–1099 (2016).

    Article  PubMed  CAS  Google Scholar 

  21. Xia, G. et al. Conformation-specific binding of p31comet antagonizes the function of Mad2 in the spindle checkpoint. EMBO J. 23, 3133–3143 (2004).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Mapelli, M. et al. Determinants of conformational dimerization of Mad2 and its inhibition by p31comet. EMBO J. 25, 1273–1284 (2006).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Mapelli, M., Massimiliano, L., Santaguida, S. & Musacchio, A. The Mad2 conformational dimer: structure and implications for the spindle assembly checkpoint. Cell 131, 730–743 (2007).

    Article  PubMed  CAS  Google Scholar 

  24. Lyubimov, A. Y., Strycharska, M. & Berger, J. M. The nuts and bolts of ring-translocase structure and mechanism. Curr. Opin. Struct. Biol. 21, 240–248 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Yang, M. et al. Insights into mad2 regulation in the spindle checkpoint revealed by the crystal structure of the symmetric mad2 dimer. PLoS Biol. 6, e50 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Luo, X. et al. The Mad2 spindle checkpoint protein has two distinct natively folded states. Nat. Struct. Mol. Biol. 11, 338–345 (2004).

    Article  PubMed  CAS  Google Scholar 

  27. Hara, M., Özkan, E., Sun, H., Yu, H. & Luo, X. Structure of an intermediate conformer of the spindle checkpoint protein Mad2. Proc. Natl Acad. Sci. USA 112, 11252–11257 (2015).

    Article  ADS  PubMed  CAS  Google Scholar 

  28. Alfieri, C. et al. Molecular basis of APC/C regulation by the spindle assembly checkpoint. Nature 536, 431–436 (2016).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  29. Zheng, S. Q. et al. MotionCor2: anisotropic correction of beam-induced motion for improved cryo-electron microscopy. Nat. Methods 14, 331–332 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Zhang, K. Gctf: Real-time CTF determination and correction. J. S. Biol. 193, 1–12 (2016).

    Article  CAS  Google Scholar 

  31. Fernandez-Leiro, R. & Scheres, S. H. W. A pipeline approach to single-particle processing in RELION. Acta Crystallogr. D Struct. Biol. 73, 496–502 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  32. Afanasyev, P. et al. Single-particle cryo-EM using alignment by classification (ABC): the structure of Lumbricus terrestris haemoglobin. IUCrJ 4, 678–694 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Chen, S. et al. High-resolution noise substitution to measure overfitting and validate resolution in 3D structure determination by single particle electron cryomicroscopy. Ultramicroscopy 135, 24–35 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  34. Yang, Z. et al. UCSF Chimera, MODELLER, and IMP: an integrated modeling system. J. Struct. Biol. 179, 269–278 (2012).

    Article  PubMed  CAS  Google Scholar 

  35. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D Biol. Crystallogr. 60, 2126–2132 (2004).

    Article  PubMed  CAS  Google Scholar 

  36. Zhang, F. et al. Structural insights into the regulatory particle of the proteasome from Methanocaldococcus jannaschii. Mol. Cell 34, 473–484 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Adams, P. D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D Biol. Crystallogr. 66, 213–221 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Landau, M. et al. ConSurf 2005: the projection of evolutionary conservation scores of residues on protein structures. Nucleic Acids Res. 33, W299–W302 (2005).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  39. Krissinel, E. & Henrick, K. Inference of macromolecular assemblies from crystalline state. J. Mol. Biol. 372, 774–797 (2007).

    Article  PubMed  CAS  Google Scholar 

  40. Waterhouse, A. M., Procter, J. B., Martin, D. M. A., Clamp, M. & Barton, G. J. Jalview Version 2—a multiple sequence alignment editor and analysis workbench. Bioinformatics 25, 1189–1191 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. Kucukelbir, A., Sigworth, F. J. & Tagare, H. D. Quantifying the local resolution of cryo-EM density maps. Nat. Methods 11, 63–65 (2014).

    Article  PubMed  CAS  Google Scholar 

  42. Chao, W. C., Kulkarni, K., Zhang, Z., Kong, E. H. & Barford, D. Structure of the mitotic checkpoint complex. Nature 484, 208–213 (2012).

    Article  ADS  PubMed  CAS  Google Scholar 

  43. Luo, X., Tang, Z., Rizo, J. & Yu, H. The Mad2 spindle checkpoint protein undergoes similar major conformational changes upon binding to either Mad1 or Cdc20. Mol. Cell 9, 59–71 (2002).

    Article  PubMed  Google Scholar 

  44. Sironi, L. et al. Crystal structure of the tetrameric Mad1-Mad2 core complex: implications of a ‘safety belt’ binding mechanism for the spindle checkpoint. EMBO J. 21, 2496–2506 (2002).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  45. Ripstein, Z. A., Huang, R., Augustyniak, R., Kay, L. E. & Rubinstein, J. L. Structure of a AAA+ unfoldase in the process of unfolding substrate. eLife. 6, https://doi.org/10.7554/eLife.25754 (2017).

  46. Monroe, N., Han, H., Shen, P. S., Sundquist, W. I. & Hill, C. P. Structural basis of protein translocation by the Vps4–Vta1 AAA ATPase. eLife. 6, https://doi.org/10.7554/eLife.24487 (2017).

  47. Gates, S. N. et al. Ratchet-like polypeptide translocation mechanism of the AAA+ disaggregase Hsp104. Science. 357, 273–279 (2017).

Download references

Acknowledgements

This work was funded by MRC and CR-UK grants to D.B. C.A. is an EMBO Advanced Fellow. We thank A. Boland for comments on the manuscript; S. Chen, C. Savva and G. McMullan for help with EM data collection; and J. Grimmett and T. Darling for computing.

Reviewer information

Nature thanks K. Corbett, H. Yu and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Author information

Authors and Affiliations

Authors

Contributions

C.A. cloned bacterially expressed p31comet, MAD2 and TRIP13 wild type and mutant constructs, purified proteins, performed the protein complex reconstitutions and biochemical analysis and mutagenesis. C.A. prepared EM grids, analysed EM data and determined the three dimensional reconstructions. C.A. collected EM data with the help of L.C. C.A. fitted coordinates and built models. D.B. directed the project. C.A. and D.B. wrote the manuscript with input from L.C.

Corresponding author

Correspondence to David Barford.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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 Biochemical characterization of MAD2-containing complexes with TRIP13.

a, TRIP13 and p31comet can extract O-MAD2 from MCC and APC/C–MCC regardless of APC15. Western blot showing the disassembly reactions together with the respective input material (i) of APC/C-bound MCC (APC/C–MCC), in either the absence (lanes 1–2) or presence of APC15 (lanes 3–6) and MCC alone (lanes 7–10) (experimental design is shown on top). Negative control reactions (lanes 3, 4 and 9, 10) were performed with the TRIP13(E253Q) mutant. MAD2 levels were detected with an anti-MAD2 antibody. Loading controls of APC4–STREP, APC15, CDC20, TRIP13 and BUB3 were detected with antibodies specific for STREP, APC15, CDC20, TRIP13 and BUB3, respectively. BUBR1 (lanes 7–10) was detected with an anti-STREP antibody. b, Western blots showing eluted size exclusion (Superdex 200 10/300 column) fractions of the MCC and MAD2 remodelling reactions catalysed by TRIP13–p31comet in the context of free MCC and APC/C–MCC. The fractions corresponding to (i) APC/C–MCC, (ii) MCC, (iii) p31comet–C-MAD2 and (iv) monomeric C-MAD2 are shown. A reference gel for size exclusion column elution fractions corresponding to p31comet–C-MAD2–CDC20–MBP, p31comet–C-MAD2 and monomeric C-MAD2 is shown in Extended Data Fig. 2a. c, Analysis of TRIP13–p31comet complexes with the MCC using size exclusion chromatography in the presence of ATP. Coomassie-stained gel showing the gel filtration fractions (chromatogram above) of p31comet–TRIP13 complexes in complex with MCC (fraction 8 is the p31comet–TRIP13 complex with C-MAD2–CDC20 and fraction 9 is the BUBR1–BUB3 complex). Input material (i) is shown on the left. d, MCC binds p31comet and not TRIP13 alone. Coomassie-stained gel showing the gel filtration fractions of the MCC (top gel) and in the presence of p31comet (middle gel) and TRIP13(E253Q) (lower gel). e, Chromatogram (top) and SDS–PAGE (bottom) of the gel filtration performed with the TRIP13(E253Q)–p31comet–C-MAD2–CDC20 complex in the presence of ATPγS. Experiments in ae were performed in triplicate with similar results. See Supplementary Fig. 1 for gel source data.

Extended Data Fig. 2 Biochemical assay for TRIP13–p31comet-catalysed O-MAD2 generation.

Shown are size exclusion (Superdex 200 10/300 column) chromatograms and corresponding Coomassie-stained gels for the Mad2 remodelling reaction catalysed by TRIP13–p31comet. a, Reference chromatograms and Coomassie-stained gels for (i) p31comet–C-MAD2–MBP–CDC20 (brown trace), (ii) p31comet–C-MAD2 (blue trace) and (iii) monomeric O-MAD2 (red trace). Chromatograms and gels are colour-coded. Monomeric O-MAD2 elutes in fractions 22–24, whereas p31comet–C-MAD2 elutes in fractions 17–19. b, Western blots for the products of the reaction of TRIP13 with (i) wild-type p31comet and MAD2 (orange trace) and (ii) wild-type p31comet and mutant C-MAD2 (C-MAD2Δ7 (seven N-terminal residues deleted)). c, Western blots for the products of the reaction with (i) mutant C-MAD2LEE, wild-type TRIP13 and p31comet (yellow trace), (ii) mutant TRIP13(E269R/D272R) and wild-type p31comet and C-MAD2 (green trace) and (iii) mutant p31comet α3–4 and wild-type TRIP13 and C-MAD2 (black trace). Experiments in ac were performed in triplicate with similar results. See Supplementary Fig. 1 for gel source data.

Extended Data Fig. 3 Cryo-EM analysis and resolution of TRIP13 complexes in this study.

a, Left, gallery of 2D class averages of TRIP13–p31–substrate showing different views representative of 50 2D classes. Right, a typical cryo-EM micrograph of TRIP13–p31–substrate representative of 3,630 micrographs. b, Fourier shell correlation (FSC) curves are shown for all the cryo-EM reconstructions determined in this study. c, Local resolution maps calculated with RESMAP41 of the TRIP13–p31–substrate complex. d, Cryo-EM density of the TRIP13–p31–substrate reconstruction shown as in Fig. 1b. e, g, Representative density quality for the ATPγS (e) and β-strand (g). In e, critical residues for the TRIP13 catalytic site are indicated: WA (Walker A), WB (Walker B), S1 (sensor 1), S2 (sensor 2) and RF (Arg finger). f, Close up of the TRIP13 pore loops interacting with C-MAD2NT from the cryo-EM density of TRIP13–p31–substrate. EM density for C-MAD2 shown in black mesh, TRIP13 in transparent coloured surface.

Extended Data Fig. 4 3D classification of TRIP13–p31–substrate full data set.

a, Local refinement (see Methods) by applying a mask covering TRIP13A/B/C/D/E monomers. b, c, 3D class averages obtained by classification (see Methods) of the TRIP13–p31–substrate full data set (b) and TRIP13-p31-substrate (c). The percentages relative to the total number of TRIP13-p31-substrate particles are shown.

Extended Data Fig. 5 Comparative analysis of TRIP13 structures.

a, Comparison of TRIP13 cryo-EM structure (right, this study) and previous TRIP13 crystal structures (left, C. elegans PCH217; middle, human TRIP1318). b, Comparison of TRIP13 monomers within the TRIP13 cryo-EM structure. RMSDs between TRIP13A and other TRIP13 monomers are indicated in the inset table. A superimposition of all six TRIP13 monomers is shown, colour-coded as in Fig. 1. TRIP13F differs from all the other monomers in the relative orientation of the small and large AAA+ domains. Its conformation relative to TRIP13A is shown at the lower right. The open conformation prevents nucleotide binding. (The sensor 2 residue (S2) is positioned too far from the ATP-binding site.) Lower left, a superimposition of TRIP13F onto an open subunit C (grey) of the PCH2 structure17. c, Conformational differences in pore loop-1 between the PCH2–ADP17 complex (grey) and the cryo-EM TRIP13–ATPγS complex (orange) (this study).

Extended Data Fig. 6 Structures of MAD2.

Structural context of MAD2. a, C-MAD223; b, in the TRIP13–p31comet–C-MAD2 complex (this work); c, O-MAD223; d, the Schizosaccharomyces pombe mitotic checkpoint complex42 composed of C-MAD2, CDC20 and BUBR1/MAD3 (BUB3 is not shown); e, p31comet–C-MAD2 complex19. In all figures the regions of MAD2 that reposition during the O-MAD2 to C-MAD2 transition are coloured blue and red for N-terminal (residues 1–16) and C-terminal (158–204) regions, respectively. In O-MAD2 these are the N terminus (MAD2NT) and β1 strand (blue) and C-terminal β7–β8 hairpin (red). In C-MAD2 these are MAD2NT including the αN helix and first turn of αA (blue), and the C-terminal β8′–β8′′ hairpin, safety belt and C terminus (MAD2CT) (red). On conversion of O-MAD2 to C-MAD2, the β1 strand is displaced and replaced by the β8′–β8′′ hairpin. Residues 13–15 of β1 form an additional turn at the N terminus of αA in C-MAD2. The C-MAD2 ligand is coloured purple. MBM, MAD2-binding motif; MBP1, high-affinity MAD2-binding peptide43,44.

Extended Data Fig. 7 TRIP13 interacts with p31comet through a composite interface formed of monomers D and E.

a, Details of the interaction between TRIP13 and p31comet. Left, schematic of TRIP13–p31comet interactions. Right, structure showing details of the main electrostatic contacts between TRIP13 and p31comet. Above, schematic of the domain architecture of TRIP13 and p31comet. A row of aspartates on α7 engages the conserved safety belt motif residues Arg233 and Arg237 of p31comet. The adjacent Lys162 contacts a Glu-rich loop (111–121) in TRIP13 that is disordered in previous TRIP13 crystal structures17,18. In our structure the Glu-rich loop lies directly above pore loop-1. Glu104 and Asp105 of the TRIP13NTD-ATPase domain linker, immediately preceding the Glu-rich loop, contact Arg227 and Lys229 of the p31comet safety-belt, agreeing with the importance of Lys229 for TRIP13-p31comet interactions in vitro17 and in vivo20. On monomer E, the same acidic patch of α7 of the large AAA+ domain contacts basic residues at the N terminus of α3 of p31comet. b, Details of the interaction of the p31comet α3–4 loop with TRIP13 subunit E. Seven basic residues shown were deleted and the mutant p31comet(α3–4 loop) was tested in MAD2 remodelling assays and for assembly of a TRIP13–p31comet–C-MAD2 complex. c, Multiple sequence alignment of the p31comet α3–4 loop. d, Deletion of the nine N-terminal residues of MAD2 (MAD2(Δ9)), and mutation of the p31comet α3–4 loop (p31comet(α3–4 loop)) do not disrupt TRIP13–p31comet–C-MAD2 complex assembly. Coomassie-stained gel showing the gel filtration fraction of wild-type and relevant mutant TRIP13–p31comet–C-MAD2 complexes purified by size exclusion chromatography. Experiment in d was performed in triplicate with similar results. See Supplementary Fig. 1 for gel source data.

Extended Data Fig. 8 Conservation of TRIP13 pore loops and MAD2NT.

ac, Multiple sequence alignment of TRIP13 pore loop-1 (a) and pore loop-2 (b) and the N-terminal region of MAD2 (c).

Extended Data Fig. 9 Models of the TRIP13–p31comet–C-MAD2 complexes in basal state 0 and basal state 1 (before and after the first catalytic cycle).

a, Basal state 0 (similar to Fig. 4a). Superscripts on TRIP13 subunit labels denote basal state. b, Showing the conformational change of the TRIP13 hexamer after one catalytic cycle (basal state 1) superimposed on the p31comet–C-MAD2 substrate before catalysis (as in basal state 0). The upward movement of the E and F subunits clashes with p31comet and C-MAD2. The pore loop residues of E and F subunits shift by 30 Å and 13 Å, respectively. c, Close up view of b showing the clash between the F1 and E1 subunits of TRIP13 in basal state 1 with C-MAD2 and p31comet, respectively (as in basal state 0). d, Basal state 1 with TRIP13 and p31comet–C-MAD2 in the remodelled conformation (as in Fig. 4b) and now with the whole of the p31comet–C-MAD2–CDC20 substrate repositioned onto the new C1 and D1 interface and the αA helix unwound by one turn. e, As in d but rotated by 60° to show the p31comet–C-MAD2–CDC20 substrate in the same view as in a. fj, Panels showing TRIP13 and only C-MAD2NT, connection to the C-MAD2 αA helix and proposed unwinding of the C-MAD2 αA helix. f, MAD2NT and αA in basal state 0 with view of TRIP13 (in basal state 1 conformation) as in e. g, TRIP13 in basal state 1 conformation as in e with the whole p31comet–C-MAD2 substrate repositioned onto the C1–D1 interface. Note that the αA helix of C-MAD2 has not been unwound and C-MAD2NT (residues 2–12) is shifted along with the globular domain of C-MAD2 (C-MAD2GD). h, Basal state 1 with C-MAD2NT in the basal state 0 position (that is, unshifted relative to basal state 0) and αA of C-MAD2GD as p31comet–C-MAD2 is repositioned on the C1–D1 interface (but the αA helix is not unwound). This shows the 11 Å break between the Cα atoms of Thr12 of C-MAD2NT and Leu13 of C-MAD2GD. C-MAD2NT is shifted down the TRIP13 pore by two residues relative to g. i, Close up view of the C-MAD2NT and connection to the αA helix. The view is the same as in h and it superimposes the positions of C-MAD2NT as in basal state 1 and basal state 0, showing a 7.6 Å distance between the Cα atoms of Thr12 of C-MAD2NT in the two states. This indicates the extent of required unwinding of the αA helix. j, The first turn of the αA helix unwinds to reconnect Thr12 of C-MAD2NT with Leu13 of the C-MAD2 αA helix. Model building was guided by insights from cryo-EM structures of the AAA+ ATPases VAT45, Vps446 and Hsp10447 that indicate a processive hand-over-hand mechanism in which the AAA+ motor translocates along the axis of the substrate. ATP hydrolysis propagates a sequential conformational change within the hexameric ring that converts the conformation of each subunit to that of its clockwise neighbour viewed from p31comet.

Extended Data Table 1 EM data collection, processing statistics and structure refinement statistics

Supplementary information

Supplementary Figure 1

Original source images for all data obtained by electrophoretic separation: Coomassie stained SDS PAGE and western blots

Reporting Summary

Video 1: Overview of the TRIP13 and the TRIP13:p31comet:C-MAD2:CDC20 complex.

This video shows an overall view of the apo TRIP13 and then the TRIP13:p31comet:C-MAD2:CDC20 complex (TRIP13-p31-substrate complex) (basal state/class 2).C- MAD2NT enters the central pore of TRIP13

Video 2: Molecular mechanism of C-MAD2 remodelling by TRIP13 in conjunction with p31comet.

Two catalytic cycles are shown. Each catalytic cycle results in the translocation of two residues of C-MAD2NT through the TRIP13 central pore in conjunction with unwinding of one turn at the N-terminus of the MAD2 αA helix. This breaks contacts to the β8’-β8” hairpin (Ser16 of αA to His191 of β8’-β8” is shown) to destabilize the C-MAD2 β-sheet, promoting the conversion of C-MAD2 to O-MAD2

Video 3: ATP hydrolysis by TRIP13 promotes a conformational change of the TRIP13 ring and a clash with p31comet:MAD2.

This video shows how ATP-dependent translocation of TRIP13 along Mad2NT causes subunits E and F to clash with p31comet and C-MAD2, respectively

Video 4: Comparison of basal and active states of the TRIP13:p31comet:C-MAD2:CDC20 complex.

This video shows conformational change between the basal and activated states of the TRIP13:p31comet:C-MAD2:CDC20 complex. The E and F monomers change conformation on transition from the basal to activated state, moving towards p31comet:C-MAD2

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Alfieri, C., Chang, L. & Barford, D. Mechanism for remodelling of the cell cycle checkpoint protein MAD2 by the ATPase TRIP13. Nature 559, 274–278 (2018). https://doi.org/10.1038/s41586-018-0281-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-018-0281-1

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