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:

DNA damage signalling targets the kinetochore to promote chromatin mobility

Abstract

In budding yeast, chromatin mobility increases after a DNA double-strand break (DSB). This increase is dependent on Mec1, the yeast ATR kinase, but the targets responsible for this phenomenon are unknown. Here we report that the Mec1-dependent phosphorylation of Cep3, a kinetochore component, is required to stimulate chromatin mobility after DNA breaks. Cep3 phosphorylation counteracts a constraint on chromosome movement imposed by the attachment of centromeres to the spindle pole body. A second constraint, imposed by the tethering of telomeres to the nuclear periphery, is also relieved after chromosome breakage. A non-phosphorylatable Cep3 mutant that impairs DSB-induced chromatin mobility is proficient in DSB repair, suggesting that break-induced chromatin mobility may be dispensable for homology search. Rather, we propose that the relief of centromeric constraint promotes cell cycle arrest and faithful chromosome segregation through the engagement of the spindle assembly checkpoint.

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

Figure 1: Relieving constraints on chromosomes mimics DSB-induced chromatin mobility.
Figure 2: Identification of a kinetochore component required for DSB-induced chromatin mobility.
Figure 3: Cep3 is a Rad53 phosphorylation target.
Figure 4: Cep3 modulates centromere attachment to the SPB.
Figure 5: Cep3 regulates the global chromatin response.
Figure 6: Centromeric constraint defines the increase in DSB-induced mobility.
Figure 7: Cep3 phosphorylation promotes cell cycle arrest and genome stability.
Figure 8: Proposed model for Cep3 following a DSB.

Similar content being viewed by others

References

  1. Marshall, W. F. et al. Interphase chromosomes undergo constrained diffusional motion in living cells. Curr. Biol. 7, 930–939 (1997).

    CAS  PubMed  Google Scholar 

  2. Chubb, J. R., Boyle, S., Perry, P. & Bickmore, W. A. Chromatin motion is constrained by association with nuclear compartments in human cells. Curr. Biol. 12, 439–445 (2002).

    CAS  PubMed  Google Scholar 

  3. Vazquez, J., Belmont, A. S. & Sedat, J. W. Multiple regimes of constrained chromosome motion are regulated in the interphase Drosophila nucleus. Curr. Biol. 11, 1227–1239 (2001).

    CAS  PubMed  Google Scholar 

  4. Heun, P., Laroche, T., Shimada, K., Furrer, P. & Gasser, S. M. Chromosome dynamics in the yeast interphase nucleus. Science 294, 2181–2186 (2001).

    CAS  PubMed  Google Scholar 

  5. Seeber, A., Dion, V. & Gasser, S. M. Checkpoint kinases and the INO80 nucleosome remodeling complex enhance global chromatin mobility in response to DNA damage. Genes Dev. 27, 1999–2008 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Chung, D. K. et al. Perinuclear tethers license telomeric DSBs for a broad kinesin- and NPC-dependent DNA repair process. Nat. Commun. 6, 7742 (2015).

    CAS  PubMed  Google Scholar 

  7. Dion, V. & Gasser, S. M. Chromatin movement in the maintenance of genome stability. Cell 152, 1355–1364 (2013).

    CAS  PubMed  Google Scholar 

  8. Krawczyk, P. M. et al. Chromatin mobility is increased at sites of DNA double-strand breaks. J. Cell Sci. 125, 2127–2133 (2012).

    CAS  PubMed  Google Scholar 

  9. Lottersberger, F., Karssemeijer, R. A., Dimitrova, N. & de Lange, T. 53BP1 and the LINC complex promote microtubule-dependent DSB mobility and DNA repair. Cell 163, 880–893 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Dimitrova, N., Chen, Y. C., Spector, D. L. & de Lange, T. 53BP1 promotes non-homologous end joining of telomeres by increasing chromatin mobility. Nature 456, 524–528 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Cho, N. W., Dilley, R. L., Lampson, M. A. & Greenberg, R. A. Interchromosomal homology searches drive directional ALT telomere movement and synapsis. Cell 159, 108–121 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Kruhlak, M. J. et al. Changes in chromatin structure and mobility in living cells at sites of DNA double-strand breaks. J. Cell Biol. 172, 823–834 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Soutoglou, E. et al. Positional stability of single double-strand breaks in mammalian cells. Nat. Cell Biol. 9, 675–682 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Roukos, V. et al. Spatial dynamics of chromosome translocations in living cells. Science 341, 660–664 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Dion, V., Kalck, V., Horigome, C., Towbin, B. D. & Gasser, S. M. Increased mobility of double-strand breaks requires Mec1, Rad9 and the homologous recombination machinery. Nat. Cell Biol. 14, 502–509 (2012).

    CAS  PubMed  Google Scholar 

  16. Mine-Hattab, J. & Rothstein, R. Increased chromosome mobility facilitates homology search during recombination. Nat. Cell Biol. 14, 510–517 (2012).

    CAS  PubMed  Google Scholar 

  17. Aylon, Y. & Kupiec, M. DSB repair: the yeast paradigm. DNA Repair 3, 797–815 (2004).

    CAS  PubMed  Google Scholar 

  18. Renkawitz, J., Lademann, C. A. & Jentsch, S. Mechanisms and principles of homology search during recombination. Nat. Rev. Mol. Cell Biol. 15, 369–383 (2014).

    CAS  PubMed  Google Scholar 

  19. Horigome, C. et al. SWR1 and INO80 chromatin remodelers contribute to DNA double-strand break perinuclear anchorage site choice. Mol. Cell 55, 626–639 (2014).

    CAS  PubMed  Google Scholar 

  20. Gasser, S. M. Visualizing chromatin dynamics in interphase nuclei. Science 296, 1412–1416 (2002).

    CAS  PubMed  Google Scholar 

  21. Verdaasdonk, J. S. et al. Centromere tethering confines chromosome domains. Mol. Cell 52, 819–831 (2013).

    CAS  PubMed  Google Scholar 

  22. Hediger, F., Neumann, F. R., Van Houwe, G., Dubrana, K. & Gasser, S. M. Live imaging of telomeres: yKu and Sir proteins define redundant telomere-anchoring pathways in yeast. Curr. Biol. 12, 2076–2089 (2002).

    CAS  PubMed  Google Scholar 

  23. Gartenberg, M. R., Neumann, F. R., Laroche, T., Blaszczyk, M. & Gasser, S. M. Sir-mediated repression can occur independently of chromosomal and subnuclear contexts. Cell 119, 955–967 (2004).

    CAS  PubMed  Google Scholar 

  24. Hill, A. & Bloom, K. Genetic manipulation of centromere function. Mol. Cell. Biol. 7, 2397–2405 (1987).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Smolka, M. B., Albuquerque, C. P., Chen, S. H. & Zhou, H. Proteome-wide identification of in vivo targets of DNA damage checkpoint kinases. Proc. Natl Acad. Sci. USA 104, 10364–10369 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Chen, S. H., Albuquerque, C. P., Liang, J., Suhandynata, R. T. & Zhou, H. A proteome-wide analysis of kinase-substrate network in the DNA damage response. J. Biol. Chem. 285, 12803–12812 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Lisby, M., Rothstein, R. & Mortensen, U. H. Rad52 forms DNA repair and recombination centers during S phase. Proc. Natl Acad. Sci. USA 98, 8276–8282 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Lisby, M., Mortensen, U. H. & Rothstein, R. Colocalization of multiple DNA double-strand breaks at a single Rad52 repair centre. Nat. Cell Biol. 5, 572–577 (2003).

    CAS  PubMed  Google Scholar 

  29. Lechner, J. & Carbon, J. A 240 kd multisubunit protein complex, CBF3, is a major component of the budding yeast centromere. Cell 64, 717–725 (1991).

    CAS  PubMed  Google Scholar 

  30. Lechner, J. A zinc finger protein, essential for chromosome segregation, constitutes a putative DNA binding subunit of the Saccharomyces cerevisiae kinetochore complex, Cbf3. EMBO J. 13, 5203–5211 (1994).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Purvis, A. & Singleton, M. R. Insights into kinetochore–DNA interactions from the structure of Cep3Delta. EMBO Rep. 9, 56–62 (2008).

    CAS  PubMed  Google Scholar 

  32. Pinsky, B. A., Kung, C., Shokat, K. M. & Biggins, S. The Ipl1-Aurora protein kinase activates the spindle checkpoint by creating unattached kinetochores. Nat. Cell Biol. 8, 78–83 (2006).

    CAS  PubMed  Google Scholar 

  33. Dorn, J. F. et al. Yeast kinetochore microtubule dynamics analyzed by high-resolution three-dimensional microscopy. Biophys. J. 89, 2835–2854 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Jeggo, P. A. & Downs, J. A. Roles of chromatin remodellers in DNA double strand break repair. Exp. Cell Res. 329, 69–77 (2014).

    CAS  PubMed  Google Scholar 

  35. Agmon, N., Liefshitz, B., Zimmer, C., Fabre, E. & Kupiec, M. Effect of nuclear architecture on the efficiency of double-strand break repair. Nat. Cell Biol. 15, 694–699 (2013).

    CAS  PubMed  Google Scholar 

  36. Rieder, C. L., Cole, R. W., Khodjakov, A. & Sluder, G. The checkpoint delaying anaphase in response to chromosome monoorientation is mediated by an inhibitory signal produced by unattached kinetochores. J. Cell Biol. 130, 941–948 (1995).

    CAS  PubMed  Google Scholar 

  37. Sanchez, Y. et al. Control of the DNA damage checkpoint by chk1 and rad53 protein kinases through distinct mechanisms. Science 286, 1166–1171 (1999).

    CAS  PubMed  Google Scholar 

  38. Wang, H., Liu, D., Wang, Y., Qin, J. & Elledge, S. J. Pds1 phosphorylation in response to DNA damage is essential for its DNA damage checkpoint function. Genes Dev. 15, 1361–1372 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Dotiwala, F., Harrison, J. C., Jain, S., Sugawara, N. & Haber, J. E. Mad2 prolongs DNA damage checkpoint arrest caused by a double-strand break via a centromere-dependent mechanism. Curr. Biol. 20, 328–332 (2010).

    CAS  PubMed  Google Scholar 

  40. Musacchio, A. & Salmon, E. D. The spindle-assembly checkpoint in space and time. Nat. Rev. Mol. Cell Biol. 8, 379–393 (2007).

    CAS  PubMed  Google Scholar 

  41. Yuen, K. W. et al. Systematic genome instability screens in yeast and their potential relevance to cancer. Proc. Natl Acad. Sci. USA 104, 3925–3930 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Warren, C. D. et al. S-phase checkpoint genes safeguard high-fidelity sister chromatid cohesion. Mol. Biol. Cell 15, 1724–1735 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Renkawitz, J., Lademann, C. A., Kalocsay, M. & Jentsch, S. Monitoring homology search during DNA double-strand break repair in vivo. Mol. Cell 50, 261–272 (2013).

    CAS  PubMed  Google Scholar 

  44. Neumann, F. R. et al. Targeted INO80 enhances subnuclear chromatin movement and ectopic homologous recombination. Genes Dev. 26, 369–383 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Chambers, A. L. et al. The INO80 chromatin remodeling complex prevents polyploidy and maintains normal chromatin structure at centromeres. Genes Dev. 26, 2590–2603 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Lee, C.-S. et al. Chromosome position determines the success of double-strand break repair. Proc. Natl Acad. Sci. USA 113, E146–E154 (2015).

    PubMed  PubMed Central  Google Scholar 

  47. Dick, A. E. & Gerlich, D. W. Kinetic framework of spindle assembly checkpoint signalling. Nat. Cell Biol. 15, 1370–1377 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Collin, P., Nashchekina, O., Walker, R. & Pines, J. The spindle assembly checkpoint works like a rheostat rather than a toggle switch. Nat. Cell Biol. 15, 1378–1385 (2013).

    CAS  PubMed  Google Scholar 

  49. Rohner, S., Gasser, S. M. & Meister, P. Modules for cloning-free chromatin tagging in Saccharomyces cerevisae. Yeast 25, 235–239 (2008).

    CAS  PubMed  Google Scholar 

  50. Sage, D., Neumann, F. R., Hediger, F., Gasser, S. M. & Unser, M. Automatic tracking of individual fluorescence particles: application to the study of chromosome dynamics. IEEE Trans. Image Process. 14, 1372–1383 (2005).

    PubMed  Google Scholar 

  51. Wybenga-Groot, L. E. et al. Structural basis of Rad53 kinase activation by dimerization and activation segment exchange. Cell Signal. 26, 1825–1836 (2014).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We are grateful to R. Szilard and members of the Durocher laboratory for critical reading of the manuscript. We also thank S. Jaspersen (Stowers Institute for Medical Research, USA), P. Hieter (University of British Columbia, Canada), G. Brown (University of Toronto, Canada), M. Meneghini (University of Toronto, Canada), R. Rothstein (Columbia University, USA), S. Gasser (Friedrich Miescher Institute, Switzerland), S. Biggins (Fred Hutchinson Cancer Research Center, USA) and M. Kupiec (Tel Aviv University, Israel) for sharing strains and plasmids. J.S. is supported by a CIHR Doctoral award. D.D. is the Thomas Kierans Chair in Mechanisms of Cancer Development and a Canada Research Chair (Tier 1) in the Molecular Mechanisms of Genome Integrity. This work was supported by CIHR grant FDN143343 (to D.D.) and MOP123468 (to L.P.) and a Grant-in-Aid from the Krembil Foundation to L.P. and D.D.

Author information

Authors and Affiliations

Authors

Contributions

J.S. carried out the experiments and wrote the manuscript. G.D.G. and M.B. generated MATLAB scripts for nuclear alignment and MSD analysis. W.Z. and M.-C.L. helped initiate the project and set up the microscopy system for measuring chromatin mobility. L.P. supervised G.D.G. and M.B. and provided microscopy advice. D.D. supervised the project and wrote the manuscript with J.S.

Corresponding author

Correspondence to Daniel Durocher.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 3 System for chromatin mobility analysis in budding yeast.

(a) Representative images of mobility strains. Chromatin is tracked by integrating LacOx256 arrays in cells expressing LacI-GFP and Nup49-mCherry. Scale bar represents 2 μm. (b) X-axis displacement of a single tracked MAT locus over time in 1 × 1 or 2 × 2 binned images. (c) Growth comparison of unexposed and imaged cells. Cells were held in a DeltaVision microscope at 30 °C in SD media supplemented with raffinose and imaged at the indicated time points. Scale bar represents 4 μm. (d) MSD analysis of the MAT locus in wild-type cells before and 3 h after DSB induction. (e) MSD analysis of the MAT locus in wild-type cells or in the INO80 mutant arp8Δ. All MSD data represent the mean ± s.e.m. of cells pooled from 3 independent experiments. See Supplementary Table 1 for mobility parameters and the precise number of cells analysed.

Supplementary Figure 4 Cep3-S575 is required for DSB-induced chromatin mobility.

(a,b) Radius of confinement (a) and diffusion coefficients (b) of wild-type and cep3-S575A cells following a DSB. Data represents the mean ± s.e.m., n represent the number of cells pooled from 3 independent experiments (WT control n = 131, WT DSB n = 81, cep3-S575A control n = 189, cep3-S575A DSB n = 111). P < 0.001; one-way ANOVA. (c) MSD analysis of the MAT locus in wild-type cells and a cep3-S575A mutant generated at the endogenous CEP3 locus by the pop-in/pop-out method. (d) MSD analysis of the MAT locus in wild-type or cep3-S575A/cep3-S575A diploid cells containing an HO cut site on one homolog. All MSD data represent the mean ± s.e.m. of cells pooled from 3 independent experiments. See Supplementary Table 1 for mobility parameters and the precise number of cells analysed.

Supplementary Figure 5 Cep3 phosphorylation requires Mec1/Rad53/Dun1 signalling.

(a) Immunoblot analysis using a Cep3-S575 phosphospecific antibody on whole cell extracts from cells expressing Myc-tagged Cep3 (WT or S575A) in response to a HO-induced DSB. Rad53 was probed to confirm checkpoint activation, Myc was used to control for loading. (b) Immunoblot analysis of asynchronous (ASN), α-factor arrested (G1, or nocodazole arrested (G2/M) cells treated with Zeocin (250 μg ml−1) for 1 h. Levels of the mitotic cyclin Clb2 were ascertained to confirm cell cycle synchronization. Pgk1 and Rad53 were probed to control for loading and to confirm checkpoint activation respectively. (c) MSD analysis of the MAT locus in sml1Δ tel1Δ mutants. (d) MSD analysis of the MAT locus in sml1Δ tel1Δ mec1Δ mutants. (e) MSD analysis of the MAT locus in cep3-S575A rad53Δ mutants. (f) Amino acid sequence surrounding Cep3-S575 and the consensus phosphorylation target sites of the Rad53 and Dun1 kinases. Matching residues are highlighted in red, Ψ represents aliphatic amino acids. (g) Time-course of Cep3-13xMyc phosphorylation in response to Zeocin (250 μg ml−1) determined by immunoblot analysis. (h) MSD analysis of the MAT locus in dun1Δ mutants. (i) Cep3-S575 phosphorylation in dun1Δ cells following 1 h of Zeocin treatment (250 μg ml−1) determined by immunoblot analysis.

Supplementary Figure 6 Mobility of a CEN-bound episome following DSB induction.

(a,b) MSD analysis of a CEN/ARS-LacOx256 episome in wild-type (a) and cep3-S575A (b) cells in response to a DSB 27 kb from CEN4. (c) MSD analysis of a CEN/ARS-LacOx256 episome in response to a DSB 998 kb from CEN4. All MSD data represent the mean ± s.e.m. of cells pooled from 3 independent experiments. See Supplementary Table 1 for mobility parameters and the precise number of cells analysed.

Supplementary Figure 7 Loss of telomeric tethering does not increase the mobility of a broken chromosome arm.

(a) MSD analysis of the MAT locus on Chr III (in cis) and the MAK10 locus on unbroken Chr V in trans to a DSB 27 kb from CEN4. (b) MSD analysis of the MAK10 locus on unbroken Chr V in trans to a DSB 27 kb from CEN4 in sir4Δ cells. (c,d,e) MSD analysis of the MAT locus, in cis to a DSB, in sir4Δ (c), yku70Δ (d), and yku70Δ sir4Δ (e) cells. All MSD data represent the mean ± s.e.m. of cells pooled from 3 independent experiments. See Supplementary Table 1 for mobility parameters and the precise number of cells analysed.

Supplementary Figure 8 Mobility of a broken and unbroken chromosome arm.

(a) Schematic of strains to follow the mobility of both chromosome arms following an HO-induced DSB on Chr IV (red triangle). (b) MSD analysis of the Chr IV-R and Chr IV-L arms before and after a DSB. (c) MSD analysis of the uncut Chr IV-L arm in wild-type and cep3-S575A cells. (d) MSD analysis of the cut Chr IV-R arm and the uncut Chr IV-L arm after a DSB in a sir4Δ background. All MSD data represent the mean ± s.e.m. of cells pooled from 3 independent experiments. See Supplementary Table 1 for mobility parameters and the precise number of cells analysed.

Supplementary Figure 9 Cep3-S75A does not affect the kinetics of DSB repair or DNA binding.

(a) Schematic of the HR repair strain NA60. PCR primers P1 and P2 were used to amplify the homologous recombination cassette containing the HO cut site on Chr IX. Repair of the DSB by HR is accompanied by the appearance of a ClaI site provided by the donor recombination cassette on Chr XIII. (b) Agarose gel of ClaI-digested PCR products in response to a DSB in wild-type and cep3-S575A cells. DNA was extracted from cells harvested at the indicated time points, the cut locus was amplified by PCR, and PCR products were digested with ClaI. Gene conversion (GC) products are cleaved by ClaI and form two lower molecular weight bands. Early repair products can be detected at 2.5 h in both wild-type and cep3-S575A cells. A colony that survived plating on galactose was used as a positive control (+). (c) Chromatin immunoprecipitation (ChIP) of Cep3-13xMyc and Cep3-S575A-13xMyc in response to a DSB at the MAT locus. Fold enrichment at CEN3 is normalized to the non-centromeric control locus TSC11. Data represents the mean ± s.d., n = 5 independent experiments for each condition.

Supplementary Figure 10 Cep3 and INO80 function in a common pathway.

(a,b) MSD analysis of the MAT locus in INO80 mutant arp8Δ cells with CEN3 (a) or CEN5 (b) inactivation. (c) Length of checkpoint arrest in response to an irreparable DSB at the MAT locus. Each dot represents a single cell combined from 2 or more independent experiments for each genotype (WT n = 199 cells, cep3-S575A n = 194, arp8Δ n = 171, cep3-S75A arp8Δ n = 171); red bars represent the median arrest length. (d) Immunoblot analysis of Cep3-S575 phosphorylation in wild-type and cep3-S575Acells, and two clones each of the INO80 mutants arp5Δ and arp8Δ after Zeocin treatment (250 μg ml−1) for 1 h (+) as described in Fig. 3c. All MSD data represent the mean ± s.e.m. of cells pooled from 3 independent experiments. See Supplementary Table 1 for mobility parameters and the precise number of cells analysed.

Supplementary Figure 11 Uncropped scans of immunoblots.

Cropped images for figures are indicated by a red box.

Supplementary Table 1 Summary of mobility parameters.
Supplementary Table 2 Strains used in this study.

Supplementary information

Supplementary Information

Supplementary Information (PDF 777 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Strecker, J., Gupta, G., Zhang, W. et al. DNA damage signalling targets the kinetochore to promote chromatin mobility. Nat Cell Biol 18, 281–290 (2016). https://doi.org/10.1038/ncb3308

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ncb3308

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