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:

The inner centromere is a biomolecular condensate scaffolded by the chromosomal passenger complex

Abstract

The inner centromere is a region on every mitotic chromosome that enables specific biochemical reactions that underlie properties, such as the maintenance of cohesion, the regulation of kinetochores and the assembly of specialized chromatin, that can resist microtubule pulling forces. The chromosomal passenger complex (CPC) is abundantly localized to the inner centromeres and it is unclear whether it is involved in non-kinase activities that contribute to the generation of these unique chromatin properties. We find that the borealin subunit of the CPC drives phase separation of the CPC in vitro at concentrations that are below those found on the inner centromere. We also provide strong evidence that the CPC exists in a phase-separated state at the inner centromere. CPC phase separation is required for its inner-centromere localization and function during mitosis. We suggest that the CPC combines phase separation, kinase and histone code-reading activities to enable the formation of a chromatin body with unique biochemical activities at the inner centromere.

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: The centromere-targeting region of the CPC undergoes phase separation in vitro under physiological conditions.
Fig. 2: The dynamics of the CPC subunits in coacervates are similar to those measured in the inner centromere.
Fig. 3: Components of the inner centromere are enriched in ISB coacervates.
Fig. 4: Inner-centromere components can induce ISB phase separation.
Fig. 5: Borealin undergoes phase separation in vivo and inner-centromeric CPC is sensitive to inhibitors of phase separation.
Fig. 6: The phase-separation property of the CPC is crucial for its localization to inner centromeres and spindle midzones.
Fig. 7: The phase-separation property of the CPC is important for its mitotic functions.

Similar content being viewed by others

Data availability

Source data for Figs. 17 and Supplementary Figs. 17 are provided in Supplementary Table 2. The data that support the findings of this study are available from the corresponding author on reasonable request.

References

  1. Trivedi, P. & Stukenberg, P. T. A centromere-signaling network underlies the coordination among mitotic events. Trends Biochem. Sci. 41, 160–174 (2016).

    Article  CAS  Google Scholar 

  2. Bloom, K. S. Centromeric heterochromatin: the primordial segregation machine. Annu. Rev. Genet. 48, 457–484 (2014).

    Article  CAS  Google Scholar 

  3. Jaqaman, K. et al. Kinetochore alignment within the metaphase plate is regulated by centromere stiffness and microtubule depolymerases. J. Cell Biol. 188, 665–679 (2010).

    Article  CAS  Google Scholar 

  4. Banani, S. F., Lee, H. O., Hyman, A. A. & Rosen, M. K. Biomolecular condensates: organizers of cellular biochemistry. Nat. Rev. Mol. Cell Biol. 18, 285–298 (2017).

    Article  CAS  Google Scholar 

  5. Larson, A. G. et al. Liquid droplet formation by HP1α suggests a role for phase separation in heterochromatin. Nature 547, 236–240 (2017).

    Article  CAS  Google Scholar 

  6. Strom, A. R. et al. Phase separation drives heterochromatin domain formation. Nature 547, 241–245 (2017).

    Article  CAS  Google Scholar 

  7. Sabari, B. R. et al. Coactivator condensation at super-enhancers links phase separation and gene control. Science 361, eaar3958 (2018).

    Article  Google Scholar 

  8. Mahen, R. et al. Comparative assessment of fluorescent transgene methods for quantitative imaging in human cells. Mol. Biol. Cell 25, 3610–3618 (2014).

    Article  Google Scholar 

  9. Sessa, F. et al. Mechanism of aurora B activation by INCENP and inhibition by hesperadin. Mol. Cell 18, 379–391 (2005).

    Article  CAS  Google Scholar 

  10. Fuller, B. G. et al. Midzone activation of aurora B in anaphase produces an intracellular phosphorylation gradient. Nature 453, 1132–1136 (2008).

    Article  CAS  Google Scholar 

  11. Wang, E., Ballister, E. R. & Lampson, M. A. Aurora B dynamics at centromeres create a diffusion-based phosphorylation gradient. J. Cell Biol. 194, 539–549 (2011).

    Article  CAS  Google Scholar 

  12. Yamagishi, Y., Honda, T., Tanno, Y. & Watanabe, Y. Two histone marks establish the inner centromere and chromosome bi-orientation. Science 330, 239–243 (2010).

    Article  CAS  Google Scholar 

  13. Wang, F. et al. Histone H3 Thr-3 phosphorylation by haspin positions aurora B at centromeres in mitosis. Science 330, 231–235 (2010).

    Article  CAS  Google Scholar 

  14. Kelly, A. E. et al. Survivin reads phosphorylated histone H3 threonine 3 to activate the mitotic kinase aurora B. Science 330, 235–239 (2010).

    Article  CAS  Google Scholar 

  15. YamagishiY., HondaT., TannoY. & WatanabeY. Two histone marks establish the inner centromere and chromosome bi-orientation. Science 330, 239–243 (2010).

    Article  CAS  Google Scholar 

  16. Tsukahara, T., Tanno, Y. & Watanabe, Y. Phosphorylation of the CPC by Cdk1 promotes chromosome bi-orientation. Nature 467, 719–723 (2010).

    Article  CAS  Google Scholar 

  17. Niedzialkowska, E. et al. Molecular basis for phosphospecific recognition of histone H3 tails by survivin paralogues at inner centromeres. Mol. Biol. Cell 23, 1457–1466 (2012).

    Article  CAS  Google Scholar 

  18. Du, J., Kelly, A. E., Funabiki, H. & Patel, D. J. Structural basis for recognition of H3T3ph and Smac/DIABLO N-terminal peptides by human survivin. Structure 20, 185–195 (2012).

    Article  CAS  Google Scholar 

  19. Delacour-Larose, M., Molla, A., Skoufias, D. A., Margolis, R. L. & Dimitrov, S. Distinct dynamics of aurora b and survivin during mitosis. Cell Cycle 3, 1418–1426 (2004).

    Article  CAS  Google Scholar 

  20. Beardmore, V. A. Survivin dynamics increases at centromeres during G2/M phase transition and is regulated by microtubule-attachment and aurora B kinase activity. J. Cell Sci. 117, 4033–4042 (2004).

    Article  CAS  Google Scholar 

  21. Wühr, M. et al. Deep proteomics of the Xenopus laevis egg using an mRNA-derived reference database. Curr. Biol. 24, 1467–1475 (2014).

    Article  Google Scholar 

  22. Hindriksen, S., Lens, S. M. A. & Hadders, M. A. The ins and outs of aurora B inner centromere localization. Front. Cell Dev. Biol. 5, 112 (2017).

    Article  Google Scholar 

  23. Hauf, S. et al. The small molecule Hesperadin reveals a role for aurora B in correcting kinetochore-microtubule attachment and in maintaining the spindle assembly checkpoint. J. Cell Biol. 161, 281–294 (2003).

    Article  CAS  Google Scholar 

  24. Hengeveld, R. C. C., Vromans, M. J. M., Vleugel, M., Hadders, M. A. & Lens, S. M. A. Inner centromere localization of the CPC maintains centromere cohesion and allows mitotic checkpoint silencing. Nat. Commun. 8, 15542 (2017).

    Article  CAS  Google Scholar 

  25. Klein, U. R., Nigg, E. A. & Gruneberg, U. Centromere targeting of the chromosomal passenger complex requires a ternary subcomplex of borealin, survivin, and the N-terminal domain of INCENP. Mol. Biol. Cell 17, 2547–2558 (2006).

    Article  CAS  Google Scholar 

  26. Woodruff, J. B. et al. The centrosome is a selective condensate that nucleates microtubules by concentrating tubulin. Cell 169, 1066–1077 (2017).

    Article  CAS  Google Scholar 

  27. Wheelock, M. S., Wynne, D. J., Tseng, B. S. & Funabiki, H. Dual recognition of chromatin and microtubules by INCENP is important for mitotic progression. J. Cell Biol. 216, 925–941 (2017).

    Article  CAS  Google Scholar 

  28. Hanley, M. L., Yoo, T. Y., Sonnett, M., Needleman, D. J. & Mitchison, T. J. Chromosomal passenger complex hydrodynamics suggests chaperoning of the inactive state by nucleoplasmin/nucleophosmin. Mol. Biol. Cell 28, 1444–1456 (2017).

    Article  CAS  Google Scholar 

  29. Wachsmuth, M. et al. High-throughput fluorescence correlation spectroscopy enables analysis of proteome dynamics in living cells. Nat. Biotechnol. 33, 384–389 (2015).

    Article  CAS  Google Scholar 

  30. Ruppert, J. G. et al. HP1α targets the chromosomal passenger complex for activation at heterochromatin before mitotic entry. EMBO J. 37, e97677 (2018).

    Article  Google Scholar 

  31. Liu, X. et al. Chromatin protein HP1α interacts with the mitotic regulator borealin protein and specifies the centromere localization of the chromosomal passenger complex. J. Biol. Chem. 289, 20638–20649 (2014).

    Article  CAS  Google Scholar 

  32. Abe, Y. et al. HP1-assisted aurora B kinase activity prevents chromosome segregation errors. Dev. Cell 36, 487–497 (2016).

    Article  CAS  Google Scholar 

  33. Chen, J. et al. Survivin enhances aurora-B kinase activity and localizes aurora-B in human cells. J. Biol. Chem. 278, 486–490 (2003).

    Article  CAS  Google Scholar 

  34. Wheatley, S. P., Carvalho, A., Vagnarelli, P. & Earnshaw, W. C. INCENP is required for proper targeting of survivin to the centromeres and the anaphase spindle during mitosis. Curr. Biol. 11, 886–890 (2001).

    Article  CAS  Google Scholar 

  35. Sampath, S. C. et al. The chromosomal passenger complex is required for chromatin-induced microtubule stabilization and spindle assembly. Cell 118, 187–202 (2004).

    Article  CAS  Google Scholar 

  36. Chakraborty, A., Prasanth, K. V. & Prasanth, S. G. Dynamic phosphorylation of HP1α regulates mitotic progression in human cells. Nat. Commun. 5, 3445 (2014).

    Article  Google Scholar 

  37. Shin, Y. et al. Spatiotemporal control of intracellular phase transitions using light-activated optoDroplets. Cell 168, 159–171 (2017).

    Article  CAS  Google Scholar 

  38. Chong, S. et al. Imaging dynamic and selective low-complexity domain interactions that control gene transcription. Science 361, eaar2555 (2018).

    Article  Google Scholar 

  39. Kroschwald, S., Maharana, S. & Simon, A. Hexanediol: a chemical probe to investigate the material properties of membrane-less compartments. Matters https://doi.org/10.19185/matters.201702000010 (2017).

  40. Jain, A. & Vale, R. D. RNA phase transitions in repeat expansion disorders. Nature 546, 243–247 (2017).

    Article  CAS  Google Scholar 

  41. Ambadipudi, S., Biernat, J., Riedel, D., Mandelkow, E. & Zweckstetter, M. Liquid–liquid phase separation of the microtubule-binding repeats of the Alzheimer-related protein Tau. Nat. Commun. 8, 275 (2017).

    Article  Google Scholar 

  42. Bolognesi, B. et al. A concentration-dependent liquid phase separation can cause toxicity upon increased protein expression. Cell Rep. 16, 222–231 (2016).

    Article  CAS  Google Scholar 

  43. Trivedi, P. et al. The binding of borealin to microtubules underlies a tension independent kinetochore-microtubule error correction pathway. Nat. Commun. 10, 682 (2019).

    Article  Google Scholar 

  44. Hirota, T., Lipp, J. J., Toh, B. H. & Peters, J. M. Histone H3 serine 10 phosphorylation by aurora B causes HP1 dissociation from heterochromatin. Nature 438, 1176–1180 (2005).

    Article  CAS  Google Scholar 

  45. Fischle, W. et al. Regulation of HP1-chromatin binding by histone H3 methylation and phosphorylation. Nature 438, 1116–1122 (2005).

    Article  CAS  Google Scholar 

  46. Ainsztein, A. M., Kandels-Lewis, S. E., Mackay, A. M. & Earnshaw, W. C. INCENP centromere and spindle targeting: identification of essential conserved motifs and involvement of heterochromatin protein HP1. J. Cell Biol. 143, 1763–1774 (1998).

    Article  CAS  Google Scholar 

  47. Rosasco-Nitcher, S. E., Lan, W., Khorasanizadeh, S. & Stukenberg, P. T. Centromeric aurora-B activation requires TD-60, microtubules, and substrate priming phosphorylation. Science 319, 469–472 (2008).

    Article  CAS  Google Scholar 

  48. Cardarelli, F., Lanzano, L. & Gratton, E. Fluorescence correlation spectroscopy of intact nuclear pore complexes. Biophys. J. 101, L27–L29 (2011).

    Article  CAS  Google Scholar 

  49. Moens, P. D. J., Gratton, E. & Salvemini, I. L. Fluorescence correlation spectroscopy, raster image correlation spectroscopy, and number and brightness on a commercial confocal laser scanning microscope with analog detectors (Nikon C1). Microsc. Res. Tech. 74, 377–388 (2011).

    Google Scholar 

  50. Scipioni, L., Lanzanó, L., Diaspro, A. & Gratton, E. Comprehensive correlation analysis for super-resolution dynamic fingerprinting of cellular compartments using the Zeiss Airyscan detector. Nat. Commun. 9, 5120 (2018).

    Article  CAS  Google Scholar 

  51. Janicki, S. M. et al. From silencing to gene expression: real-time analysis in single cells. Cell 116, 683–698 (2004).

    Article  CAS  Google Scholar 

  52. Banerjee, B., Kestner, C. A. & Stukenberg, P. T. EB1 enables spindle microtubules to regulate centromeric recruitment of aurora B. J. Cell Biol. 204, 947–963 (2014).

    Article  CAS  Google Scholar 

  53. Sathyan, K. M., Fachinetti, D. & Foltz, D. R. α-Amino trimethylation of CENP-A by NRMT is required for full recruitment of the centromere. Nat. Commun. 8, 14678 (2017).

    Article  Google Scholar 

Download references

Acknowledgements

We thank J. Ellenberg, C. Brangwynne, G. Narlikar and D. Foltz for reagents, and D. Burke and S. Mattada for discussions. P.T.S. and P.T. were funded by NIH grant numbers R01GM124042 and R24OD023697. F.P. and M.A.D. were supported in part by a grant from the NSF (MCB-1615701). M.A.D. and E.G. were funded by NIH grant number P41-GM103540.

Author information

Authors and Affiliations

Authors

Contributions

P.T. and P.T.S. conceived and designed the study. P.T. performed and analysed all of the experiments, except for those involving FCS, under the guidance of P.T.S. F.P. performed and analysed the FCS experiments under the guidance of M.A.D. and E.G. E.N. provided the human pHP1α and X. laevis ISD proteins. P.T. and P.T.S. wrote the manuscript.

Corresponding author

Correspondence to P. Todd Stukenberg.

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 ISB-WT and mutant protein complexes.

(a) Superdex-200 gel filtration profile of ISBWT (red), ISBΔ139-160 (green), and ISBΔ165-180 (blue) broken grey line shows the profile of the FPLC standard. Brown bar indicated the fractions from gel filtration run on SDS-PAGE and stained with Coomassie brilliant blue (b). For a and b, the experiment was performed once for each protein prep. Source data for a are shown in Supplementary Table 2.

Supplementary Figure 2 Phase-separation properties of the CPC centromere-targeting region.

(a) DIC micrographs of the Xenopus. laevis INCENP1-58-Survivin-Dasra (ISD) complex coacervates under indicated conditions. 3 independent repeats. (b) Circularity of the GFP-ISB coacervates from experiment shown in Fig. 1d. n=257 and n=369 coacervates for 50 mM NaCl and 5% PEG respectively. Red line represents mean of the data. (c) Surface wetting and shearing of the ISB coacervates (orange arrow) under the indicated conditions. (d) Micrographs showing ISB coacervates under increasing NaCl concentration. (e) Graph showing area of the coacervates from d under increasing NaCl concentration (n=60, 71, 98 and 91 coacervates were analyzed for 50, 75, 100 and 125 mM NaCl respectively). (f) Micrographs showing ISB coacervates under increasing ISB concentration. (g) Graph showing area of the coacervates from F under increasing concentration of the ISB (n=74, 89, 58 and 98 coacervates were analyzed for 5,11, 21 and 42 μM ISB respectively). (h) Micrographs of ISB coacervates formed in presence of increasing concentration of PEG-3350. (i) Graph showing area of coacervates under increasing concentration of PEG-3350 (n=105, 182 and 149 coacervates were analyzed for 2,3 and 5% PEG respectively). Combined data from 3 independent experiments was reported for e,g and i. n.s. P>0.05, *** P<0.0001. Box and whisker graphs: median (central line), 25th-75th percentile (box), and 5th-95th percentile (whiskers). For statistical analysis in e, g, and i Kruskal-Wallis test was applied followed by Dunn’s multiple comparison test. (j and k) Electron micrographs of ISB coacervates formed under indicated conditions showing highly circular coacervates and coacervates arrest mid-fusion due to gelation. A few apparently aggregated networks could also be observed under condition of low salt and are shown as contrast to spherical coacervates. (l) DIC micrographs of 20 μM ISB showing coacervates arrest mid-fusion (yellow arrow). (m) Control from GFP-ISB FRAP analysis (Fig. 1f and Fig. 2a, b) showing no significant bleaching of GFP-ISB coacervates over the duration of imaging. For c-m all experiments were repeated 3 times independently. Scale bar 5 μm except for j and k where it is 1 μm and for the middle panel in k it is 2 μm. Source data for b, e, g and i are shown in Supplementary Table 2.

Supplementary Figure 3 FCS measurements of GFP-Borealin in the cytoplasm and at the inner-centromere.

(a) Equations used to fit the FCS correlation curves. b) The residuals of FCS fitted curves for GFP-Borealin in the cytoplasm of mitotic cell and at the inner-centromere (c) of the mitosis. (d) Graph showing diffusion coefficients of the fast population of GFP-Borealin inside the cytoplasm (Grey) or inner-centromere (Blue) during mitosis (n=25 regions were analyzed in both inner-centromere and cytoplasm). n.s. indicates P= 0.37. (e) Graph of Gslow/Gfast for GFP-Borealin in the cytoplasm (Grey) or inner-centromere (Blue) of a mitotic cell (n=25 regions were analyzed in both inner-centromere and cytoplasm). *** indicates P= 2.38E-04. For d and e statistical analysis was performed by applying two-tailed unpaired Mann-Whitney test. (f) FCS Autocorrelation functions of Inner centromere populations DFast (black line), DSlow (red line) and FCS autocorrelation curve in coacervates (green line). For b-f combined data from three independent repeats. For the whole figure same data as shown in Fig. 2b was analyzed. Box and whisker graphs: median (central line), 25th-75th percentile (box), and 5th-95th percentile (whiskers). Source data for b-f are shown in Supplementary Table 2.

Supplementary Figure 4 Enrichment of macromolecules in ISB phase and nucleation of microtubules from the ISB coacervates.

Representative images and graph showing partition coefficient of (a) Histone H3.3 mono-nucleosome (n=98), (b) Cy3-azide (n=33) and (c) GFP-xMad2 (n=101) in ISB phase. (d) Control images for experiment show in Fig. 3a. (e) Representative images and graph showing partition coefficient of rhodamine-α/β-tubulin dimer (n=111). (f) Images showing nucleation of microtubules upon enrichment of Cy5-α/β-tubulin dimers in ISB coacervates. Nucleation was observed from ISB coacervates at 1.25 uM α/β-tubulin dimers concentration in presence of GTP and not seen in the absence of GTP, the experiment was repeated independently twice. Yellow arrow points to the microtubules. “n” is the number of coacervates analyzed per condition. Box and whisker graphs: median (central line), 25th-75th percentile (box), and 5th-95th percentile (whiskers). Scale bar 5 μm. Source data for a-c and e are shown in Supplementary Table 2.

Supplementary Figure 5 Effect of increased monovalent cation on stability of ISB coacervates and Cat-granule profiles of CPC components.

(a) Graph showing enrichment of mCherry signal in the foci at 95 s. Data from experiment shown in Fig. 5a (n=99 puncta (11 cells) for BorWT-mch-cry2 and n=42 puncta (11 cells) for Bor139-160-mch-cry2; 17 cells were analyzed for mch-cry2 alone; the experiment was repeated twice). Middle brown line indicates mean and error bars indicate s.d.. For statistical analysis two tailed unpaired T-test with Welch’s correction was used, *** P<0.0001. (b) Graph showing average DIC (top) and mCherry (bottom) intensities along the line across a lacO arrays from experiment shown in Fig. 5b (n=22 LacO spots for mCherry-LacI-Borealin and n=23 LacO spot for mCherry-LacI; the experiment was repeated twice). Error-bars are s.e.m., the black dotted line in DIC graph is a reference line showing a flat profile. (Top right) insets from Fig. 5b with cyan arrow showing an example of line and its direction along which the intensities were measured and plotted. (c) DIC micrographs under indicated conditions showing disruption of ISB coacervates by 3% 1,6-Hexanediol (same conditions as in Fig. 5c, d). (d) DIC micrographs under indicated conditions showing disruption of ISB coacervates by 90 mM NH4OAC (same conditions as in Fig. 5e, f). (e) DIC micrographs of ISB coacervates in buffer containing indicated amount of NaCl and PEG-3350 (same condition as in Fig. 5g, h). For c-e, the experiment was repeated twice. (f) Graph showing normalized fluorescent intensity of Aurora-B-GFP at the inner-centromeres under indicated conditions from experiment shown in Fig. 5e (n=318, 430, 412 and 892 inner-centromeres were analyzed for 25 mM NaCl, 220 mM NaCl, 25 mM NaCl + 10% PEG-3350, and 220 mM + 10% PEG-3350 respectively). For statistical analysis two-tailed T-test with Welch’s corrections was applied, ***P<0.0001. Box and whisker graphs: median (central line), 25th-75th percentile (box), and 5th-95th percentile (whiskers). Experiment in f was repeated twice independently. (g) Catgranule profile of Borealin, Survivin and INCENP showing high propensity of Borealin central domain to participate in phase-separation. Orange and brown shaded region shows 139-160 and 165-180 amino acids of Borealin respectively. (h) PONDR profile of Borealin showing disorder propensity of the protein, Orange and brown shaded region shows 139-160 and 165-180 amino acids of Borealin respectively. (i) Alignment of Borealin region predicted to drive granule formation. Orange and brown box indicate the putative regions diving granule formation. Red and blue arrows indicate residues mutated in Borealin8A and Borealin13A mutants respectively. Scale bar is 5 μm. Source data for a, b and f are shown in Supplementary Table 2.

Supplementary Figure 6 Phase-separation property of the CPC is important for its inner-centromere localization.

(a and b) Western blots showing endogenous Borealin, LAP-Borealin and Tubulin levels. Cells were treated as indicated in Fig. 6f, except for treatment with 3.3 μM Nocodazole. Lysates for western blotting were collected 8 hours after second thymidine release. (c) Western blot showing relative levels of endogenous Borealin to LAP-BorealinWT or LAP-BorealinΔ139-160, cells were treated with 3.3 μM Nocodazole and arrested for 20 hours to enrich for mitotic cells, followed by cell lysis and western blot. (d) Micrographs showing staining of INCENP and ACA in cells rescued with either LAP-BorealinWT or LAP-BorealinΔ139-160. Cells were treated as shown in Fig. 6f, with the exception of the nocodazole treatment. For a-d the experiment was performed once. (e) Graph showing normalized intensity (normalized to that in cells expressing LAP-BorealinWT) of INCENP/ACA from experiment shown in d. (n=172 and n=157 inner-centromeres for LAP-BorealinWT and LAP-BorealinΔ139-160 respectively). Micrographs showing staining of H3T3ph (f), H2ApT120 (h), hSgo1 (j) and ACA in cells rescued with either LAP-BorealinWT or LAP-BorealinΔ139-160; cells were treated as shown in Fig. 6f. For f,h and j experiment was repeated twice. Graph showing normalized intensity (normalized to that in cells expressing LAP-BorealinWT) of H3T3ph (n=210 inner-centromeres from 10 cells for LAP-BorealinWT, n=170 inner-centromeres from 8 cells for LAP-BorealinΔ139-160, P=0.0005) (g), H2AT120ph (n=214 centromeres from 10 cells for LAP-BorealinWT, n=172 centromeres from 10 cells for LAP-BorealinΔ139-160, P=0.3494), (i), and hSgo1 (n=173 centromeres from 7 cells for LAP-BorealinWT, n=153 centromeres from 8 cells for LAP-BorealinΔ139-160, P=0.3094), (k) from experiment shown in f,h,and j respectively. (l) Micrographs of ISBWT and ISBΔ139-160 coacervates formed under indicated condition showing enrichment of CF555-hSgo11-112. Experiment repeated twice. (m) Graph showing partition coefficient of CF555-hSgo11-112 in ISBWT (n=104 coacervates) and ISBΔ139-160 (n=66 coacervates) from experiment shown in l, combined data from two independent experiments, P=0.092. For statistical analysis in e,g,i and k two-tailed T-test was applied with Welch’s correction, n.s. P>0.05, *** P<0.0001. All box and whisker graphs represent the median (central line), 25th-75th percentile (bounds of the box), and 5th-95th percentile (whiskers). Scale bar is 5 μm. For a-e experiment was done once. For f-k experiment was repeated twice. Source data for e, g, i, k and m are shown in Supplementary Table 2. Unprocessed blots for a-c are shown in Supplementary Fig. 8.

Supplementary Figure 7 Phase-separation by the CPC is important for maintenance of taxol induced mitotic arrest but only marginally affects the time to align chromosomes.

Cumulative frequency plot of the duration of Nuclear envelop break down (NEBD) to metaphase (a), metaphase to anaphase (b), and NEBD to anaphase (c); data from experiment shown in Fig. 7a, b. For statistical analysis two-tailed T-test with Welch’s correction was applied. (d) Cumulative frequency plot of the duration of mitotic arrest in presence of 100 nM paclitaxel from Fig. 7d–f. For statistical analysis two-tailed Mann-Whitney test was applied. Mean ± SEM is shown in parenthesis and “n” indicating the number of cells analyzed. Source data for a-d are shown in Supplementary Table 2.

Supplementary Figure 8 Uncropped Blots.

(a-c) Uncropped blots from the indicated figure. Dotted box indicates the cropped area shown in the figure.

Supplementary information

Supplementary Information

Supplementary Figures 1–8 and their legends, legends for Supplementary Tables 1 and 2.

Reporting Summary

Supplementary Table 1

Supplementary Table 2

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Trivedi, P., Palomba, F., Niedzialkowska, E. et al. The inner centromere is a biomolecular condensate scaffolded by the chromosomal passenger complex. Nat Cell Biol 21, 1127–1137 (2019). https://doi.org/10.1038/s41556-019-0376-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-019-0376-4

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