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:

Dynamic kinetochore size regulation promotes microtubule capture and chromosome biorientation in mitosis

Abstract

Faithful chromosome segregation depends on the ability of sister kinetochores to attach to spindle microtubules. The outer layer of kinetochores transiently expands in early mitosis to form a fibrous corona, and compacts following microtubule capture. Here we show that the dynein adaptor Spindly and the RZZ (ROD–Zwilch–ZW10) complex drive kinetochore expansion in a dynein-independent manner. C-terminal farnesylation and MPS1 kinase activity cause conformational changes of Spindly that promote oligomerization of RZZ-Spindly complexes into a filamentous meshwork in cells and in vitro. Concurrent with kinetochore expansion, Spindly potentiates kinetochore compaction by recruiting dynein via three conserved short linear motifs. Expanded kinetochores unable to compact engage in extensive, long-lived lateral microtubule interactions that persist to metaphase, and result in merotelic attachments and chromosome segregation errors in anaphase. Thus, dynamic kinetochore size regulation in mitosis is coordinated by a single, Spindly-based mechanism that promotes initial microtubule capture and subsequent correct maturation of attachments.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Spindly recruits dynein to compact kinetochores after microtubule attachment.
Fig. 2: Kinetochores expand by forming a structurally stable kinetochore sub-module.
Fig. 3: Spindly and RZZ are essential for kinetochore expansion.
Fig. 4: Spindly stimulates RZZ-Spindly polymerization in vitro and in vivo.
Fig. 5: A structural conformation of Spindly prevents RZZS oligomerization.
Fig. 6: Release of Spindly autoinhibition promotes its interaction with RZZ.
Fig. 7: MPS1 promotes RZZS meshwork formation and kinetochore expansion.
Fig. 8: The expanded kinetochore module interacts with microtubule lattices and prevents biorientation.

Similar content being viewed by others

References

  1. Musacchio, A. & Desai, A. A molecular view of kinetochore assembly and function. Biology 6, 5 (2017).

    Google Scholar 

  2. Joglekar, A. P. & Kukreja, A. A. How kinetochore architecture shapes the mechanisms of its function. Curr. Biol. 27, R816–R824 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Etemad, B. & Kops, G. J. Attachment issues: kinetochore transformations and spindle checkpoint silencing. Curr. Opin. Cell Biol. 39, 101–108 (2016).

    CAS  PubMed  Google Scholar 

  4. Maiato, H. The dynamic kinetochore-microtubule interface. J. Cell Sci. 117, 5461–5477 (2004).

    CAS  PubMed  Google Scholar 

  5. Maiato, H., Gomes, A., Sousa, F. & Barisic, M. Mechanisms of chromosome congression during mitosis. Biology 6, 13 (2017).

    Google Scholar 

  6. Magidson, V. et al. Adaptive changes in the kinetochore architecture facilitate proper spindle assembly. Nat. Cell Biol. 17, 1134–1144 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Wynne, D. J. & Funabiki, H. Heterogeneous architecture of vertebrate kinetochores revealed by three-dimensional superresolution fluorescence microscopy. Mol. Biol. Cell 27, 3395–3404 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Thrower, D. A., Jordan, M. A. & Wilson, L. Modulation of CENP-E organization at kinetochores by spindle microtubule attachment. Cell Motil. Cytoskelet. 35, 121–133 (1996).

    CAS  Google Scholar 

  9. Hoffman, D. B., Pearson, C. G., Yen, T. J., Howell, B. J. & Salmon, E. D. Microtubule-dependent changes in assembly of microtubule motor proteins and mitotic spindle checkpoint proteins at PtK1 kinetochores. Mol. Biol. Cell 12, 1995–2009 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Magidson, V. et al. Unattached kinetochores rather than intrakinetochore tension arrest mitosis in taxol-treated cells. J. Cell Biol. 212, 307–319 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Wynne, D. J. & Funabiki, H. Kinetochore function is controlled by a phosphodependent coexpansion of inner and outer components. J. Cell Biol. 210, 899–916 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Rieder, C. L. The formation, structure, and composition of the mammalian kinetochore and kinetochore fiber. Int. Rev. Cytol. 79, 1–58 (1982).

    CAS  PubMed  Google Scholar 

  13. McEwen, B. F., Hsieh, C.-E., Mattheyses, A. L. & Rieder, C. L. A new look at kineochore structure in vertebrate somatic cells using high-pressure freezing an freeze substitution. Chromosoma 107, 366–375 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Jokelainen, P. T. The ultrastructure and spatial organization of the metaphase kinetochore in mitotic rat cells. J. Ultrastruct. Res. 19, 19–44 (1967).

    CAS  PubMed  Google Scholar 

  15. Cassimeris, L., Rieder, C. L., Rupp, G. & Salmon, E. D. Stability of microtubule attachment to metaphase kinetochores in PtK1 cells. J. Cell Sci. 96, 9–15 (1990). Pt 1.

    PubMed  Google Scholar 

  16. McEwen, B. F., Arena, J. T., Frank, J. & Rieder, C. L. Structure of the Colcemid-treated PtK1 kinetochore outer plate as determined by high voltage electron microscopic tomography. J. Cell Biol. 120, 301–312 (1993).

    CAS  PubMed  Google Scholar 

  17. Pereira, A. L. et al. Mammalian CLASP1 and CLASP2 cooperate to ensure mitotic fidelity by regulating spindle and kinetochore function. Mol. Biol. Cell 17, 4526–42 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Wordeman, L., Steuer, E. R., Sheetz, M. P. & Mitchison, T. Chemical subdomains within the kinetochore domain of isolated CHO mitotic chromosomes. J. Cell Biol. 114, 285–94 (1991).

    CAS  PubMed  Google Scholar 

  19. Yao, X., Anderson, K. L. & Cleveland, D. W. The microtubule-dependent motor centromere-associated protein E (CENP- E) is an integral component of kinetochore corona fibers that link centromeres to spindle microtubules. J. Cell Biol. 139, 435–447 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Cooke, C. A., Schaar, B., Yen, T. J. & Earnshaw, W. C. Localization of CENP-E in the fibrous corona and outer plate of mammalian kinetochores from prometaphase through anaphase. Chromosoma 106, 446–455 (1997).

    CAS  PubMed  Google Scholar 

  21. Mosalaganti, S. et al. Structure of the RZZ complex and molecular basis of its interaction with Spindly. J. Cell Biol. 216, 961–981 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Kops, G. J. P. L. et al. ZW10 links mitotic checkpoint signaling to the structural kinetochore. J. Cell Biol. 169, 49–60 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Buffin, E., Lefebvre, C., Huang, J., Gagou, M. E. & Karess, R. E. Recruitment of Mad2 to the kinetochore requires the Rod/Zw10 complex. Curr. Biol. 15, 856–861 (2005).

    CAS  PubMed  Google Scholar 

  24. Défachelles, L. et al. RZZ and Mad1 dynamics in Drosophila mitosis. Chromosom. Res. 23, 333–342 (2015).

    Google Scholar 

  25. Starr, D. A. et al. Conservation of the centromere/kinetochore protein ZW10. J. Cell Biol. 138, 1289–1301 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Gama, J. B. et al. Molecular mechanism of dynein recruitment to kinetochores by the Rod-Zw10-Zwilch complex and Spindly. J. Cell Biol. 216, 943–960 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Çivril, F. et al. Structural analysis of the RZZ complex reveals common ancestry with multisubunit vesicle tethering machinery. Structure 18, 616–626 (2010).

    PubMed  Google Scholar 

  28. Basto, R. et al. In vivo dynamics of the rough deal checkpoint protein during Drosophila mitosis. Curr. Biol. 14, 56–61 (2004).

    CAS  PubMed  Google Scholar 

  29. Howell, B. J. et al. Cytoplasmic dynein/dynactin drives kinetochore protein transport to the spindle poles and has a role in mitotic spindle checkpoint inactivation. J. Cell Biol. 155, 1159–1172 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Wojcik, E. et al. Kinetochore dynein: its dynamics and role in the transport of the Rough deal checkpoint protein. Nat. Cell Biol. 3, 1001–1007 (2001).

    CAS  PubMed  Google Scholar 

  31. Famulski, J. K., Vos, L. J., Rattner, J. B. & Chan, G. K. Dynein/dynactin-mediated transport of kinetochore components off kinetochores and onto spindle poles induced by Nordihydroguaiaretic acid. PLoS One 6, e16494 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Griffis, E. R., Stuurman, N. & Vale, R. D. Spindly, a novel protein essential for silencing the spindle assembly checkpoint, recruits dynein to the kinetochore. J. Cell Biol. 177, 1005–1015 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Gassmann, R. et al. A new mechanism controlling kinetochore-microtubule interactions revealed by comparison of two dynein-targeting components: SPDL-1 and the Rod/Zwilch/Zw10 complex. Genes Dev. 22, 2385–2399 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Ying, W. C. et al. Mitotic control of kinetochore-associated dynein and spindle orientation by human Spindly. J. Cell Biol. 185, 859–874 (2009).

    Google Scholar 

  35. McKenney, R. J., Huynh, W., Tanenbaum, M. E., Bhabha, G. & Vale, R. D. Activation of cytoplasmic dynein motility by dynactin-cargo adapter complexes. Science 345, 337–341 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Moudgil, D. K. et al. A novel role of farnesylation in targeting a mitotic checkpoint protein, human spindly, to kinetochores. J. Cell Biol. 208, 881–896 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Holland, A. J. et al. Preventing farnesylation of the dynein adaptor Spindly contributes to the mitotic defects caused by farnesyltransferase inhibitors. Mol. Biol. Cell 26, 1845–1856 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Barisic, M. et al. Spindly/CCDC99 is required for efficient chromosome congression and mitotic checkpoint regulation. Mol. Biol. Cell 21, 1968–1981 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Gassmann, R. et al. Removal of Spindly from microtubule-attached kinetochores controls spindle checkpoint silencing in human cells. Genes Dev. 24, 957–971 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Tromer, E., Bade, D., Snel, B. & Kops, G. J. P. L. Phylogenomics-guided discovery of a novel conserved cassette of short linear motifs in BubR1 essential for the spindle checkpoint. Open Biol. 6, 1–11 (2016).

    Google Scholar 

  41. Njoroge, F. G. et al. (+)-4-[2-[4-(8-Chloro-3,10-dibromo-6,11-dihydro-5H-benzo[5,6]cyclohepta[1,2-b]- pyridin-11(R)-yl)-1-piperidinyl]-2-oxo-ethyl]-1-piperidinecarboxamide (SCH-66336): a very potent farnesyl protein transferase inhibitor as a novel antitumor agent. J. Med. Chem. 41, 4890–4902 (1998).

    CAS  PubMed  Google Scholar 

  42. Nijenhuis, W. et al. A TPR domain-containing N-terminal module of MPS1 is required for its kinetochore localization by Aurora B. J. Cell Biol. 201, 217–231 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Brauchle, M. et al. Protein interference applications in cellular and developmental biology using DARPins that recognize GFP and mCherry. Biol. Open 3, 1252–1261 (2014).

    PubMed  PubMed Central  Google Scholar 

  44. Rieder, C. L. & Alexander, S. P. Kinetochores are transported poleward along a single astral microtubule during chromosome attachment to the spindle in newt lung cells. J. Cell Biol. 110, 81–95 (1990).

    CAS  PubMed  Google Scholar 

  45. Kapoor, T. M. et al. Chromosomes can congress to the metaphase plate before biorientation. Science 311, 388–91 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Shrestha, R. L. & Draviam, V. M. Lateral to end-on conversion of chromosome-microtubule attachment requires kinesins cenp-e and MCAK. Curr. Biol. 23, 1514–1526 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Kajtez, J. et al. Overlap microtubules link sister k-fibres and balance the forces on bi-oriented kinetochores. Nat. Commun. 7, 10298 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Chozinski, T. J. et al. Expansion microscopy with conventional antibodies and fluorescent proteins. Nat. Methods 13, 485–488 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Gregan, J., Polakova, S., Zhang, L., Tolić-Nørrelykke, I. M. & Cimini, D. Merotelic kinetochore attachment: causes and effects. Trends Cell Biol. 21, 374–381 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Pesenti, M. E., Weir, J. R. & Musacchio, A. Progress in the structural and functional characterization of kinetochores. Curr. Opin. Struct. Biol. 37, 152–163 (2016).

    CAS  PubMed  Google Scholar 

  51. Hoogenraad, C. C. & Akhmanova, A. Bicaudal D family of motor adaptors: linking dynein motility to cargo binding. Trends Cell Biol. 26, 327–340 (2016).

    CAS  PubMed  Google Scholar 

  52. Ghosal, D. et al. Collaborative protein filaments. EMBO J. 34, 2312–2320 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Zhang, G., Lischetti, T., Hayward, D. G. & Nilsson, J. Distinct domains in Bub1 localize RZZ and BubR1 to kinetochores to regulate the checkpoint. Nat. Commun. 6, 7162 (2015).

    CAS  PubMed  Google Scholar 

  54. Caldas, G. V. et al. The RZZ complex requires the N-terminus of KNL1 to mediate optimal Mad1 kinetochore localization in human cells. Open Biol. 5, 150160 (2015).

    PubMed  PubMed Central  Google Scholar 

  55. Karess, R. Rod-Zw10-Zwilch: a key player in the spindle checkpoint. Trends Cell Biol. 15, 386–392 (2005).

    CAS  PubMed  Google Scholar 

  56. Steigemann, P. et al. Aurora B-mediated abscission checkpoint protects against tetraploidization. Cell 136, 473–484 (2009).

    PubMed  Google Scholar 

  57. Saurin, A. T., van der Waal, M. S., Medema, R. H., Lens, S. M. A. & Kops, G. J. P. L. Aurora B potentiates Mps1 activation to ensure rapid checkpoint establishment at the onset of mitosis. Nat. Commun. 2, 316 (2011).

    PubMed  Google Scholar 

  58. Dosztanyi, Z., Csizmok, V., Tompa, P. & Simon, I. The pairwise energy content estimated from amino acid composition discriminates between folded and intrinsically unstructured proteins. J. Mol. Biol. 347, 827–839 (2005).

    CAS  PubMed  Google Scholar 

  59. Delorenzi, M. & Speed, T. An HMM model for coiled-coil domains and a comparison with PSSM-based predictions. Bioinformatics 18, 617–625 (2002).

    CAS  PubMed  Google Scholar 

  60. Bailey, T. L. et al. MEME SUITE: tools for motif discovery and searching. Nucleic Acids Res. 37, W202–208 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Katoh, K. & Standley, D. M. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol. Biol. Evol. 30, 772–780 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Eddy, S. R. Accelerated profile HMM searches. PLoS Comput. Biol. 7, e1002195 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Weissmann, F. et al. biGBac enables rapid gene assembly for the expression of large multisubunit protein complexes. Proc. Natl Acad. Sci. USA 113, 2564–2569 (2016).

    Google Scholar 

  64. de la Rosa-Trevín, J. M. et al. Scipion: a software framework toward integration, reproducibility and validation in 3D electron microscopy. J. Struct. Biol. 195, 93–99 (2016).

    PubMed  Google Scholar 

  65. Abrishami, V. et al. A pattern matching approach to the automatic selection of particles from low-contrast electron micrographs. Bioinformatics 29, 2460–2468 (2013).

    CAS  PubMed  Google Scholar 

  66. Sorzano, C. O. S. et al. A clustering approach to multireference alignment of single-particle projections in electron microscopy. J. Struct. Biol. 171, 197–206 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Pernot, P. et al. Upgraded ESRF BM29 beamline for SAXS on macromolecules in solution. J. Synchrotron Radiat. 20, 660–664 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Petoukhov, M. V. et al. New developments in the ATSAS program package for small-angle scattering data analysis. J. Appl. Crystallogr. 45, 342–350 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Rambo, R. P. & Tainer, J. A. Accurate assessment of mass, models and resolution by small-angle scattering. Nature 496, 477–481 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Iyer, L. M., Zhang, D., Maxwell Burroughs, A. & Aravind, L. Computational identification of novel biochemical systems involved in oxidation, glycosylation and other complex modifications of bases in DNA. Nucleic Acids Res. 41, 7635–7655 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank all lab members for suggestions and discussions. We are grateful to A. Murachelli for help with EM data figure preparation; to E. von Castelmur, T. Heidebrecht and Y. Hiruma for help with Spindly structure experiments; to J. Vaughan for help with ExM; to R. Gassmann for sharing unpublished results and Spindly constructs; to I. Cheeseman, S. Lens and R. Medema for reagents; and to A. de Graaf of the Hubrecht Imaging Center. The Horizon 2020 iNEXT project (653706) provided financial support and access to EM infrastructures. This work is part of the Oncode Institute which is partly financed by the Dutch Cancer Society. This work was further supported by the Netherlands Organisation for Scientific Research (NWO) (gravitation program CancerGenomiCs.nl; VICI grant (865.12.004 to G.J.P.L.K.)), the Dutch Cancer Society (KWF/HUBR-11080 to G.J.P.L.K.), and the ERC (675737 to A.M.). V.G. is supported by the Proteins@Work initiative of the Netherlands Proteomics Centre.

Author information

Authors and Affiliations

Authors

Contributions

C.S. and G.J.P.L.K. conceived the project. C.S., G.J.P.L.K., M.U.D.A., A.P., J.K. and A.M designed experiments and interpreted data. C.S. performed the cell biology experiments. M.U.D.A. and J.K. performed the in vitro experiments with the help of A.F. J.F. and J.K. performed and analysed the electron microscopy experiments. V.G. performed the cross-linking experiments. E.T. performed the comparative sequence analysis. R.M. and J.M.C performed the electron microscopy of Spindly. C.S. and G.J.P.L.K. wrote the manuscript with the help of A.P. and A.M. and the input of the rest of authors.

Corresponding author

Correspondence to Geert JPL Kops.

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.

Integrated supplementary information

Supplementary Figure 1 Inhibition of mitotic kinases is not sufficient to compact expanded kinetochores. Related to Figure 1.

(a-b) Scheme of workflow (a) and representative images (b) of immunostains of nocodazole-treated HeLa cells incubated with inhibitors of Aurora B (ZM-447439), PLK1 (BI-2636) and MPS1 (Cpd-5). The experiment was repeated at least three times with similar results.

Supplementary Figure 2 Multiple sequence alignment of dynein adaptors. Related to Figure 1.

Multiple sequence alignment of the CC1, CC2 and Spindly box motifs found in a selection of metazoan orthologs of the human spindly-like dynein-dynactin adaptors SPINDLY, TRAK1/2, HAP1 BICD1/2 and BICDR1/2 (supplementary sequences 2). Metazoan species: Homo sapiens (human), Mus musculus (mouse), Xenopus tropicalis (frog), Danionrerio (zebra fish), Drosophila melanogaster (fruit fly), Caenorhabditis elegans (worm), Nematostella vectensis (starlet sea anemone). Sequences used can be found in Supplementary Tables 4 and 5.

Supplementary Figure 3 Characterization of dynein-binding motifs of Spindly. Related to Figure 1.

(a) Immunoblot showing Spindly knockdown and expression of siRNA-resistant GFP-tagged versions of Spindly in mitotic population. The experiment was repeated twice with similar results. (b) Representative immunofluorescence images of HeLa cells transfected with siRNA to Spindly and treated or not with doxycycline (Dox) to induce the expression of GFP-Spindly variants, and immunostained for the indicated antigens. The experiment was repeated at least three times with similar results. The same images of FL and SpindlyΔN are shown in Figure 1f. (c) Representative images of nocodazole-treated HeLa cells transfected with siRNAs to Spindly and expressing the indicated GFP-Spindly variants, and immunostained for the indicated antigens. Quantification is shown in Figure 1c. (d) Representative images of metaphase HeLa cells transfected with siRNA to Spindly and expressing the indicated GFP-Spindly variants, and immunostained for the indicated antigens. Quantification is shown in Figure 1e. The same images of FL, SpindlyΔSB and SpindlyΔCCS are shown in Figure 1f. FL, Full length Spindly.

Supplementary Figure 4 Characterization of kinetochore expansion. Related to Figure 3.

(a) Cartoon depicting the effect of ZW10 or Spindly depletion on the expandable module of the kinetochore (green), KMN network (blue) and CCAN (red). (b-c) Immunostaining with the indicated antibodies (b) and quantification (c) of kinetochore levels in cells from one of the experiments shown in Figure 3e,f. The graph in (c) shows the mean kinetochore intensity normalized to the values of the control. Control is siGAPDH. See Supplementary Table 3 for source data.

Supplementary Figure 5 Polymerization of the RZZ complex by SpindlyΔN. Related to Figure 4.

(a) Example of a HeLa cell expressing GFP-SpindlyΔN and showing a cytoplasmic filament positive for Zwilch and ROD. The experiment was repeated at least three times with similar results.

Supplementary Figure 6 Spindly protein analysis. Related to Figure 5.

(a) Coomassie staining of SDS-PAGE gels of the constructs used in Figure 5. (b) Volume of correlation plot from SEC-SAXS analysis for Spindly1-440 and SpindlyFL. Guinier extrapolated dataset plotted as q.I(q) (left panel) and integrated area of q (right panel). The calculated Vc and Rg values for Spindly1-440 are 1096 Å2 and 76Å, respectively. For, SpindlyFL, the Vc and Rg values are 1562 Å2 and 88 Å, respectively. The molecular mass69 calculated is 128 KDa and 220 KDa for Spindly1-440 and SpindlyFL, respectively. See Supplementary Table 3 for source data. (c) SEC-MALLS analysis of Spindly1-440 and SpindlyFL. The absolute molar mass in daltons (log scale) is plotted against elution volume. Calculated molar masses corresponding to the gel filtration peaks are shown as light (Spindly1-440) or dark blue (SpindlyFL) curves. For comparison, the expected molar masses for a monomer, dimer and a trimer are also plotted. The analysis shows that Spindly1-440 is dimeric in solution whereas in similar conditions SpindlyFL exists as a trimer. The experiment was repeated twice with similar results. See Supplementary Table 3 for source data. (d) Cross-linking proteomics analysis of SpindlyFL. All the cross-linked residues (boxes in grey) found in two independent experiments are represented in the X and Y axis. All the lysines present in Spindly are depicted in red. The encircled areas (b,c,d) mark several long-range interactions between the N-terminal region (1-250) and the lysines encircled in purple, suggesting intramolecular interactions between the N-terminal domain and the middle region of Spindly. See Supplementary Table 3 for source data.

Supplementary Figure 7 Release of Spindly intramolecular interactions enhances the association of Spindly with RZZ. Related to Figure 6.

(a) Binding of RZZ to the indicated versions of Spindly constructs isolated from insect cells determined by Surface Plasmon Resonance (SPR). The response plotted on the y axis is normalized for the molecular weight of the analyte to yield the stoichiometry of binding. The experiment was repeated three times with similar results. See Supplementary Table 3 for source data. (b) Immunoblot showing GFP-Spindly versions in mitotic cells treated as indicated (Lon, Lonafarnib). The experiment was repeated twice with similar results. (c-f) Representative images (c,d) and quantification (e,f) of relative kinetochore intensities of the indicated GFP-Spindly variants in nocodazole-treated HeLa cells transfected with siRNA to Spindly. The graphs show the mean kinetochore intensity normalized to the values of SpindlyFL. Each dot represents one cell. The sample size in (e) is: FL (n= 54 cells), FL/C602A (n= 55 cells), ΔN (n= 56 cells), ΔN/C602A (n= 56 cells), Δ274 (n= 51 cells) Δ274/C602A (n= 55 cells), Δ287 (n= 55 cells), Δ287/C602A (n= 51 cells) pooled from two independent experiments. See Supplementary Table 3 for source data. The sample size in (f) is: FL (n= 63 cells), FL/C602A (n= 57 cells), ΔCCS (n= 61 cells), ΔCCS/C602A (n= 59 cells) pooled from two independent experiments. See Supplementary Table 3 for source data.

Supplementary Figure 8 Enlarged kinetochores in metaphase have immature attachments. Related to Figure 8.

(a,b) Representative images (a) and quantification of kinetochore levels (b) of Astrin in metaphase cells transfected with siRNA to Spindly and expressing the indicated GFP-Spindly variants. The graph shows the mean kinetochore intensity normalized to SpindlyFL. Each dot represents one cell: FL (n=55 cells), ΔN (n=48 cells) pooled from two independent experiments. Source data can be found in Supplementary Table 3. (c) Representative images of metaphase HeLa cells transfected with siRNA to Spindly, expressing GFP-SpindlyΔN and mCherry-Tubulin, fixed at 37°C or after cold treatment, and stained for indicated antigens. (‘b’, bridging fibre; ‘k’, k-fibres). The asterisk indicates the region of the enlarged kinetochore that normally mediates the lateral interaction with microtubules at 37°C. The experiment was repeated at least three times with similar results. (d) Live-cell imaging of HeLa cells transfected with siRNA to Spindly and expressing GFP-SpindlyΔN and mCherry-Tubulin. (‘Gr’, Growing microtubules; ‘Sh’, Shrinking microtubules; ‘e’ end-on attachment; ‘l’, lateral attachment). Maximum projections of several z-planes are shown. See also Supplementary Movies 5. The experiment was repeated at least three times with similar results. (e) Model of the mechanism of kinetochore expansion and compaction performed by the axis RZZ-Spindly-dynein and MPS1. See discussion for details.

Supplementary information

Supplementary Information

Supplementary Figures 1–8, Supplementary Table and Supplementary Video legends

Reporting Summary

Supplementary Table 1

List of plasmids, primers and RNAi sequences used in this study

Supplementary Table 2

List of antibodies used in this study

Supplementary Table 3

Statistics Source Data

Supplementary Table 4

Related to Figure 1b. All sequences of Spindly used in this study

Supplementary Table 5

Related to Supplementary Figure 2. All sequences used in this study

Supplementary Video 1

Formation of filaments in vitro by RZZS

Supplementary Video 2

Formation of cytoplasmic filaments in cells expressing GFP-SpindlyΔN

Supplementary Video 3

Interaction of expanded kinetochores with microtubules

Supplementary Video 4

Interaction of expanded kinetochores with microtubules

Supplementary Video 5

Interaction of expanded kinetochores with microtubules

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sacristan, C., Ahmad, M.U.D., Keller, J. et al. Dynamic kinetochore size regulation promotes microtubule capture and chromosome biorientation in mitosis. Nat Cell Biol 20, 800–810 (2018). https://doi.org/10.1038/s41556-018-0130-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-018-0130-3

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing