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:

Microtubule-associated proteins control the kinetics of microtubule nucleation

Subjects

Abstract

Microtubules are born and reborn continuously, even during quiescence. These polymers are nucleated from templates, namely γ-tubulin ring complexes (γ-TuRCs) and severed microtubule ends. Using single-molecule biophysics, we show that nucleation from γ-TuRCs, axonemes and seed microtubules requires tubulin concentrations that lie well above the critical concentration. We measured considerable time lags between the arrival of tubulin and the onset of steady-state elongation. Microtubule-associated proteins (MAPs) alter these time lags. Catastrophe factors (MCAK and EB1) inhibited nucleation, whereas a polymerase (XMAP215) and an anti-catastrophe factor (TPX2) promoted nucleation. We observed similar phenomena in cells. We conclude that GTP hydrolysis inhibits microtubule nucleation by destabilizing the nascent plus ends required for persistent elongation. Our results explain how MAPs establish the spatial and temporal profile of microtubule nucleation.

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: Microtubule nucleation from templates is a sigmoid function of tubulin concentration.
Figure 2: Microtubule plus ends elongate at tubulin concentrations where templates do not.
Figure 3: There is a time lag associated with templated nucleation.
Figure 4: Catastrophe factors slow down templated nucleation.
Figure 5: TPX2 is an anti-catastrophe factor that accelerates nucleation.
Figure 6: A polymerase and GMPCPP–tubulin both accelerate nucleation.
Figure 7: Depletion of tubulin by nocodazole reduces the rate of nucleation in cells.

Similar content being viewed by others

References

  1. Voter, W. A. & Erickson, H. P. The kinetics of microtubule assembly. Evidence for a two-stage nucleation mechanism. J. Biol. Chem. 259, 10430–10438 (1984).

    CAS  PubMed  Google Scholar 

  2. Zheng, Y., Wong, M. L., Alberts, B. & Mitchison, T. Nucleation of microtubule assembly by a γ-tubulin-containing ring complex. Nature 378, 578–583 (1995).

    Article  CAS  PubMed  Google Scholar 

  3. Moritz, M., Braunfeld, M. B., Sedat, J. W., Alberts, B. & Agard, D. A. Microtubule nucleation by γ-tubulin-containing rings in the centrosome. Nature 378, 638–640 (1995).

    Article  CAS  PubMed  Google Scholar 

  4. Srayko, M., Kaya, A., Stamford, J. & Hyman, A. A. Identification and characterization of factors required for microtubule growth and nucleation in the early C. elegans embryo. Dev. Cell. 9, 223–236 (2005).

    Article  CAS  PubMed  Google Scholar 

  5. McNally, K., Audhya, A., Oegema, K. & McNally, F. J. Katanin controls mitotic and meiotic spindle length. J. Cell Biol. 175, 881–891 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Lindeboom, J. J. et al. A mechanism for reorientation of cortical microtubule arrays driven by microtubule severing. Science 342, 1245533 (2013).

    Article  CAS  PubMed  Google Scholar 

  7. Kollman, J. M., Polka, J. K., Zelter, A., Davis, T. N. & Agard, D. A. Microtubule nucleating γ-TuSC assembles structures with 13-fold microtubule-like symmetry. Nature 466, 879–882 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Kollman, J. M., Merdes, A., Mourey, L. & Agard, D. A. Microtubule nucleation by γ-tubulin complexes. Nat. Rev. Mol. Cell Biol. 12, 709–721 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Oakley, B. R., Oakley, C. E., Yoon, Y. & Jung, M. K. γ-tubulin is a component of the spindle pole body that is essential for microtubule function in Aspergillus nidulans. Cell 61, 1289–1301 (1990).

    Article  CAS  PubMed  Google Scholar 

  10. Luders, J. & Stearns, T. Microtubule-organizing centres: a re-evaluation. Nat. Rev. Mol. Cell Biol. 8, 161–167 (2007).

    Article  CAS  PubMed  Google Scholar 

  11. Baas, P. W. & Ahmad, F. J. The plus ends of stable microtubules are the exclusive nucleating structures for microtubules in the axon. J. Cell Biol. 116, 1231–1241 (1992).

    Article  CAS  PubMed  Google Scholar 

  12. Piehl, M., Tulu, U. S., Wadsworth, P. & Cassimeris, L. Centrosome maturation: measurement of microtubule nucleation throughout the cell cycle by using GFP-tagged EB1. Proc. Natl Acad. Sci. USA 101, 1584–1588 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Brugues, J., Nuzzo, V., Mazur, E. & Needleman, D. J. Nucleation and transport organize microtubules in metaphase spindles. Cell 149, 554–564 (2012).

    Article  CAS  PubMed  Google Scholar 

  14. Luders, J., Patel, U. K. & Stearns, T. GCP-WD is a γ-tubulin targeting factor required for centrosomal and chromatin-mediated microtubule nucleation. Nat. Cell Biol. 8, 137–147 (2006).

    Article  CAS  PubMed  Google Scholar 

  15. Bre, M. H. & Karsenti, E. Effects of brain microtubule-associated proteins on microtubule dynamics and the nucleating activity of centrosomes. Cell Motil. Cytoskeleton 15, 88–98 (1990).

    Article  CAS  PubMed  Google Scholar 

  16. Gard, D. L. & Kirschner, M. W. A microtubule-associated protein from Xenopus eggs that specifically promotes assembly at the plus-end. J. Cell Biol. 105, 2203–2215 (1987).

    Article  CAS  PubMed  Google Scholar 

  17. Brouhard, G. J. et al. XMAP215 is a processive microtubule polymerase. Cell 132, 79–88 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Popov, A. V., Severin, F. & Karsenti, E. XMAP215 is required for the microtubule-nucleating activity of centrosomes. Curr. Biol. 12, 1326–1330 (2002).

    Article  CAS  PubMed  Google Scholar 

  19. Reber, S. B. et al. XMAP215 activity sets spindle length by controlling the total mass of spindle microtubules. Nat. Cell Biol. 15, 1116–1122 (2013).

    Article  CAS  PubMed  Google Scholar 

  20. Belmont, L. D. & Mitchison, T. J. Identification of a protein that interacts with tubulin dimers and increases the catastrophe rate of microtubules. Cell 84, 623–631 (1996).

    Article  CAS  PubMed  Google Scholar 

  21. Ringhoff, D. N. & Cassimeris, L. Stathmin regulates centrosomal nucleation of microtubules and tubulin dimer/polymer partitioning. Mol. Biol. Cell 20, 3451–3458 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Heald, R. et al. Self-organization of microtubules into bipolar spindles around artificial chromosomes in Xenopus egg extracts. Nature 382, 420–425 (1996).

    Article  CAS  PubMed  Google Scholar 

  23. Carazo-Salas, R. E. et al. Generation of GTP-bound Ran by RCC1 is required for chromatin-induced mitotic spindle formation. Nature 400, 178–181 (1999).

    Article  CAS  PubMed  Google Scholar 

  24. Wittmann, T., Boleti, H., Antony, C., Karsenti, E. & Vernos, I. Localization of the kinesin-like protein Xklp2 to spindle poles requires a leucine zipper, a microtubule-associated protein, and dynein. J. Cell Biol. 143, 673–685 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Gruss, O. J. et al. Chromosome-induced microtubule assembly mediated by TPX2 is required for spindle formation in HeLa cells. Nat. Cell Biol. 4, 871–879 (2002).

    Article  CAS  PubMed  Google Scholar 

  26. Goshima, G., Mayer, M., Zhang, N., Stuurman, N. & Vale, R. D. Augmin: a protein complex required for centrosome-independent microtubule generation within the spindle. J. Cell Biol. 181, 421–429 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Petry, S., Groen, A. C., Ishihara, K., Mitchison, T. J. & Vale, R. D. Branching microtubule nucleation in Xenopus egg extracts mediated by augmin and TPX2. Cell 152, 768–777 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Schatz, C. A. et al. Importin α-regulated nucleation of microtubules by TPX2. EMBO J. 22, 2060–2070 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Mitchison, T. & Kirschner, M. Microtubule assembly nucleated by isolated centrosomes. Nature 312, 232–237 (1984).

    Article  CAS  PubMed  Google Scholar 

  30. Walker, R. A. et al. Dynamic instability of individual microtubules analyzed by video light microscopy: rate constants and transition frequencies. J. Cell Biol. 107, 1437–1448 (1988).

    Article  CAS  PubMed  Google Scholar 

  31. Mitchison, T. J. Localization of an exchangeable GTP binding site at the plus end of microtubules. Science 261, 1044–1047 (1993).

    Article  CAS  PubMed  Google Scholar 

  32. Meurer-Grob, P., Kasparian, J. & Wade, R. H. Microtubule structure at improved resolution. Biochemistry 40, 8000–8008 (2001).

    Article  CAS  PubMed  Google Scholar 

  33. Moritz, M., Braunfeld, M. B., Guenebaut, V., Heuser, J. & Agard, D. A. Structure of the γ-tubulin ring complex: a template for microtubule nucleation. Nat. Cell Biol. 2, 365–370 (2000).

    Article  CAS  PubMed  Google Scholar 

  34. Gardner, M. K. et al. Rapid microtubule self-assembly kinetics. Cell 146, 582–592 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Oosawa, F. & Asakura, S. Thermodynamics of the Polymerization of Protein Vol. 20 (Academic Press, 1975).

    Google Scholar 

  36. Mitchison, T. & Kirschner, M. Dynamic instability of microtubule growth. Nature 312, 237–242 (1984).

    Article  CAS  PubMed  Google Scholar 

  37. Drechsel, D. N., Hyman, A. A., Cobb, M. H. & Kirschner, M. W. Modulation of the dynamic instability of tubulin assembly by the microtubule-associated protein tau. Mol. Biol. Cell 3, 1141–1154 (1992).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Hunter, A. W. et al. The kinesin-related protein MCAK is a microtubule depolymerase that forms an ATP-hydrolyzing complex at microtubule ends. Mol. Cell 11, 445–457 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Gardner, M. K., Zanic, M., Gell, C., Bormuth, V. & Howard, J. Depolymerizing kinesins Kip3 and MCAK shape cellular microtubule architecture by differential control of catastrophe. Cell 147, 1092–1103 (2011).

    Article  CAS  PubMed  Google Scholar 

  40. Helenius, J., Brouhard, G., Kalaidzidis, Y., Diez, S. & Howard, J. The depolymerizing kinesin MCAK uses lattice diffusion to rapidly target microtubule ends. Nature 441, 115–119 (2006).

    Article  CAS  PubMed  Google Scholar 

  41. Askham, J. M., Vaughan, K. T., Goodson, H. V. & Morrison, E. E. Evidence that an interaction between EB1 and p150(Glued) is required for the formation and maintenance of a radial microtubule array anchored at the centrosome. Mol. Biol. Cell 13, 3627–3645 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Bieling, P. et al. Reconstitution of a microtubule plus-end tracking system in vitro. Nature 450, 1100–1105 (2007).

    Article  CAS  PubMed  Google Scholar 

  43. Vitre, B. et al. EB1 regulates microtubule dynamics and tubulin sheet closure in vitro. Nat. Cell Biol. 10, 415–421 (2008).

    Article  CAS  PubMed  Google Scholar 

  44. Gruss, O. J. et al. Ran induces spindle assembly by reversing the inhibitory effect of importin α on TPX2 activity. Cell 104, 83–93 (2001).

    Article  CAS  PubMed  Google Scholar 

  45. Brunet, S. et al. Characterization of the TPX2 domains involved in microtubule nucleation and spindle assembly in Xenopus egg extracts. Mol. Biol. Cell 15, 5318–5328 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Hyman, A. A., Salser, S., Drechsel, D. N., Unwin, N. & Mitchison, T. J. Role of GTP hydrolysis in microtubule dynamics: information from a slowly hydrolyzable analogue, GMPCPP. Mol. Biol. Cell 3, 1155–1167 (1992).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Hoebeke, J., Nijen, G. V. & Brabander, M. D. Interaction of oncodazole (r 17934), a new anti-tumoral drug, with rat brain tubulin. Biochem. Biophys. Res. Commun. 69, 319–324 (1976).

    Article  CAS  PubMed  Google Scholar 

  48. Head, J., Lee, L. L., Field, D. J. & Lee, J. C. Equilibrium and rapid kinetic studies on nocodazole-tubulin interaction. J. Biol. Chem. 260, 11060–11066 (1985).

    CAS  PubMed  Google Scholar 

  49. Vasquez, R. J., Howell, B., Yvon, A. M., Wadsworth, P. & Cassimeris, L. Nanomolar concentrations of nocodazole alter microtubule dynamic instability in vivo and in vitro. Mol. Biol. Cell 8, 973–985 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Andrews, P. D. et al. Aurora B regulates MCAK at the mitotic centromere. Dev. Cell. 6, 253–268 (2004).

    Article  CAS  PubMed  Google Scholar 

  51. Vinh, D. B., Kern, J. W., Hancock, W. O., Howard, J. & Davis, T. N. Reconstitution and characterization of budding yeast γ-tubulin complex. Mol. Biol. Cell 13, 1144–1157 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Kollman, J. M. et al. The structure of the γ-tubulin small complex: implications of its architecture and flexibility for microtubule nucleation. Mol. Biol. Cell 19, 207–215 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Kollman, J. M. et al. Ring closure activates yeast γ-turc for species-specific microtubule nucleation. Nat. Struct. Mol. Biol. 22, 132–137 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Choi, Y. K., Liu, P., Sze, S. K., Dai, C. & Qi, R. Z. CDK5RAP2 stimulates microtubule nucleation by the γ-tubulin ring complex. J. Cell Biol. 191, 1089–1095 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Chretien, D., Fuller, S. D. & Karsenti, E. Structure of growing microtubule ends: two-dimensional sheets close into tubes at variable rates. J. Cell Biol. 129, 1311–1328 (1995).

    Article  CAS  PubMed  Google Scholar 

  56. Bechstedt, S., Lu, K. & Brouhard, G. J. Doublecortin recognizes the longitudinal curvature of the microtubule end and lattice. Curr. Biol. 24, 2366–2375 (2014).

    Article  CAS  PubMed  Google Scholar 

  57. Wang, H. W. & Nogales, E. Nucleotide-dependent bending flexibility of tubulin regulates microtubule assembly. Nature 435, 911–915 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Rice, L. M., Montabana, E. A. & Agard, D. A. The lattice as allosteric effector: structural studies of αβ- and γ-tubulin clarify the role of GTP in microtubule assembly. Proc. Natl Acad. Sci. USA 105, 5378–5383 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Janosi, I. M., Chretien, D. & Flyvbjerg, H. Modeling elastic properties of microtubule tips and walls. Eur. Biophys. J. 27, 501–513 (1998).

    Article  CAS  PubMed  Google Scholar 

  60. VanBuren, V., Cassimeris, L. & Odde, D. J. Mechanochemical model of microtubule structure and self-assembly kinetics. Biophys. J. 89, 2911–2926 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Castoldi, M. & Popov, A. V. Purification of brain tubulin through two cycles of polymerization-depolymerization in a high-molarity buffer. Protein Expr. Purif. 32, 83–88 (2003).

    Article  CAS  PubMed  Google Scholar 

  62. Wieczorek, M., Chaaban, S. & Brouhard, G. Macromolecular crowding pushes catalyzed microtubule growth to near the theoretical limit. Cell. Mol. Bioeng. 6, 383–392 (2013).

    Article  Google Scholar 

  63. Gell, C. et al. Microtubule dynamics reconstituted in vitro and imaged by single-molecule fluorescence microscopy. Methods Cell Biol. 95, 221–245 (2010).

    Article  CAS  PubMed  Google Scholar 

  64. Waterman-Storer, C. M. Microtubule/organelle motility assays. Curr. Protoc. Cell Biol. Unit 13.1 (2001).

  65. Bechstedt, S. & Brouhard, G. J. Doublecortin recognizes the 13-protofilament microtubule cooperatively and tracks microtubule ends. Dev. Cell. 23, 181–192 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Norholm, M. A mutant Pfu DNA polymerase designed for advanced uracil-excision DNA engineering. BMC Biotechnol. 10, 21 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Schneider, C. A., Rasband, W. S. & Eliceiri, K. W. NIH Image to ImageJ: 25 years of image analysis. Nat. Methods 9, 671–675 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Demchouk, A. O., Gardner, M. K. & Odde, D. J. Microtubule tip tracking and tip structures at the nanometer scale using digital fluorescence microscopy. Cell. Mol. Bioeng. 4, 192–204 (2011).

    Article  PubMed  Google Scholar 

  69. Noujaim, M., Bechstedt, S., Wieczorek, M. & Brouhard, G. J. Microtubules accelerate the kinase activity of aurora-b by a reduction in dimensionality. PLoS ONE 9, e86786 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Rusan, N. M., Fagerstrom, C. J., Yvon, A. M. & Wadsworth, P. Cell cycle-dependent changes in microtubule dynamics in living cells expressing green fluorescent protein-α tubulin. Mol. Biol. Cell 12, 971–980 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank H. Higgs for suggesting the hysteresis experiment (Fig. 2) during the question period of an American Society for Cell Biology mini-symposium. We thank L. Cassimeris (Lehigh University, USA) for providing the LLCPK1:EB1–GFP cell line, A. Bird (Max Planck Institute of Molecular Physiology, Germany) for providing the U2OS:EB3–mCherry cell line, and R. Ohi (Vanderbilt University, USA) for providing the LLCPK1:GFP–tubulin cell line. We thank the Cell Imaging and Analysis Network for microscopy support. We thank C. Rocha for help with immunoblot sample preparation. We thank A. Bird, C. Friel, J. Howard, L. Rice and M. Zanic for comments on the manuscript. This work was supported by the Canadian Institutes of Health Research (CIHR, MOP-111265 to G.J.B.), by the Natural Sciences and Engineering Research Council of Canada (NSERC, no. 372593-09 to G.J.B.), and by McGill University. M.W. is supported by an NSERC CGS-D award and a Fonds de Recherche du Quebec—Nature et Technologie Bourse de Doctorat en Recherche. S.C. is supported by the NSERC CREATE training program in the Cellular Dynamics of Macromolecular Complexes and an NSERC CGS-D. G. Brouhard is a CIHR New Investigator.

Author information

Authors and Affiliations

Authors

Contributions

M.W., S.B. and G.J.B. conceived the project. M.W. performed nucleation experiments, characterized the MAPs, and imaged cells. S.B. established infrastructure for molecular biology and protein expression. S.C. performed electron microscopy experiments. M.W. and G.J.B. analysed data, developed models, and wrote the paper.

Corresponding author

Correspondence to Gary J. Brouhard.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Controls for centrosome, axoneme and GMPCPP nucleation experiments in Fig. 1.

(A) Immunoblot against γ-tubulin for a purified centrosome fraction. (B) Immunofluorescence of purified centrosomes after microtubule nucleation showing αβ-tubulin (left) and γ-tubulin (middle). In the merged image (right), the γ-tubulin signal is at the centre of the microtubule aster. (C) Overlay of our centrosome nucleation data (Fig. 1c) onto data from Fig. 4A of ref. 29, NPG. (D) Overlay of our axoneme nucleation data (Fig. 1f) onto data from Fig. 6 of ref. 30, Rockefeller Univ. Press. (E) Images from two consecutive nucleation experiments performed on the same set of GMPCPP seeds. In experiment #1, 33% of the seeds produced microtubules (left, white arrows). In experiment #2, after thorough rinsing, 25% of the seeds produced microtubules (right, white arrows). The theoretical probability that a seed produced microtubules twice agrees well with the measured value (see text at right). n = 214 GMPCPP seeds measured consecutively in the same experiment.

Supplementary Figure 2 Controls for hysteresis experiments in Fig. 2 and in-house spontaneous nucleation measurements.

(A) Plot of the average background intensity from fluorescent tubulin against time for the experiment in Fig. 2d. The intensity increases when the tubulin concentration is increased from 4 μM to 15 μM. When the solution is exchanged from 15 μM back to 4 μM, the intensity returns to the baseline, 4 μM tubulin level, indicating that the solution exchange is complete. Error bars represent the s.d. of intensity values from a 20 × 20 pixel box from one experiment. (B) Plot of the absorbance at 350 nm (A350), or turbidity, against time for solutions containing the indicated concentration of tubulin. An increase in A350 indicates polymer formation. (C) Plot of the A350 values at t = 60 min against tubulin concentration. A fit to data for which A350 > 0.05 (red line) gives an x-intercept of 21 ± 4.2 μM tubulin, which is the critical concentration for spontaneous nucleation. (D) Images of Coomassie-stained SDS-PAGE gels for the spin-down spontaneous nucleation assay, showing the total tubulin (top gel) and the polymeric tubulin in the pellet (bottom gel). (E) Plot of the concentration of pelleted tubulin against total tubulin concentration. A fit of the data for which the concentration of tubulin in the pellet ≥ 1 μM (red line) gives an x-intercept of 21 ± 7.9 μM. For increasing tubulin concentrations, n = 2, 4, 5, 5, 5, 5, 5, 5, 5, 4 and 2 independent experiments, respectively.

Supplementary Figure 3 Controls for nucleation time distribution experiments in Fig. 3.

(A) Plot of the first nucleation time against the second nucleation time for two consecutive experiments in which the same set of GMPCPP seeds were exposed to 12 μM tubulin for t = 15 min. If a seed did not produce a microtubule during the experiment, its nucleation time was recorded as >15 min. A red line shows the result predicted by the hypothesis that the seeds will have identical nucleation times in both experiments. The data clearly do not fall on the line. n = 118 GMPCPP seeds. (B) Plot of the background intensity from fluorescent tubulin against time averaged from several experiments. The intensity increases quickly as tubulin is introduced. The red line shows a fit of the data to an exponential function, y(t) = y0 + aekt. From this fit, the time at which the intensity reaches 95% of its plateau was calculated (t95% = 4 s, black dotted line). Error bars represent s.d. n = 3 independent flow-in experiments. (C) Plot of the flow cell temperature against time during a typical experiment. The red line is a fit of the temperature data to an exponential function, y(t) = y0 + aekt. From this fit, the time at which the temperature reaches 95% of its plateau was calculated (t95% = 17 s, black dotted line). Error bars represent s.d. n = 3 independent experiments.

Supplementary Figure 4 Controls for the effects of MCAK and EB1 on nucleation in Fig. 4.

(A) Image of a Coomassie-stained SDS-PAGE gel showing the purified protein fractions for MCAK, EB1, and EB1-GFP used in this study. (B) Plot of the depolymerization rate of GMPCPP microtubules against MCAK concentration. Error bars represent the s.d. For increasing MCAK concentrations, n = 10, 11, 10, 11 and 7 GMPCPP microtubules analysed in one experiment. (C) Kymographs showing double-cycled GMPCPP seeds used in our nucleation assays without (left) and with (right) 10 nM MCAK. Double-cycled GMPCPP seeds are resistant to depolymerization at these MCAK concentrations. (D) Cumulative frequency distribution of the time until catastrophe in the presence of 200 nM EB1 (red) and in control buffers (blue) at 10 μM tubulin. The solid lines are fits to the Gamma distribution, as described in Gardner et al.(2011). n = 186 (with 200 nM EB1) and n = 111 (without EB1) catastrophe events collected from different experiments. (E) In the absence of added salt, EB1-GFP binds along the GMPCPP seeds, the GDP lattice and the tip of growing microtubules. Adding 100 mM KCl to the imaging solution reduces the affinity of EB1-GFP to the seed and lattice, but end-binding persists. The tubulin concentration in these experiments was 20 μM. The EB1-GFP concentration in these experiments was 200 nM. (F) In high salt conditions, EB1 still makes nucleation difficult (green squares), arguing that EB1 lattice binding does not contribute to nucleation inhibition. The solid green line is the sigmoidal equation fit. The fit from the control data is shown in light blue for comparison. For increasing tubulin concentrations, n = 102, 92, 105 and 140 GMPCPP seeds pooled from 3 experiments. Error bars represent s.e.m.

Supplementary Figure 5 Controls for the effects of TPX2, XMAP215 and GMPCPP–tubulin on nucleation in Figs 5 and 6.

(A) Image of a Coomassie-stained SDS-PAGE gel of a microtubule cosedimenation assay done with 200 nM TPX2. The control lane indicates a calculated amount of TPX2 that would represent 100% cosedimentation. The solid triangle points to the cropped area of the original gel provided in Supplementary Fig. 7B. (B) Image of a Coomassie-stained SDS-PAGE gel of the SUMO-tagged TPX2 truncation constructs NT, CT1 and CT2 expressed and purified from bacteria. The solid triangle points to the cropped area of the original gel provided in Supplementary Fig. 7C. (C) (Top) Image of a Coomassie-stained SDS-PAGE gel of a cosedimentation assay done with 200 nM of NT, CT1 or CT2 in the presence (+) and absence (−) of 1 μM taxol-stabilized microtubules. (Bottom) Image of a Western blot against the 6xHis-tag confirming that only NT cosediments with microtubules. (D) Image of a Coomassie-stained SDS-PAGE gel showing the purified protein fractions for XMAP215 and rKin430-GFP. The solid triangle points to the cropped area of the original gel shown in Supplementary Fig. 7D. (E) Plot of microtubule growth rates against tubulin concentration in the presence of 200 nM XMAP215 (red) and in control buffers (blue). For increasing tubulin concentrations, n = 3, 29, 41 and 53 microtubules, respectively with 200 nM XMAP215 and 31, 47, 21 and 18 microtubules, respectively without XMAP215. Data were pooled across 3 experiments. Error bars represent s.e.m. (F) Plot of the nucleation probability in the presence of 200 nM rKin430-GFP (kinesin-1). The solid red line is the sigmoidal equation fit. The fit from the control data is shown in light blue for comparison. For increasing tubulin concentrations, n = 15, 10, 10, 30, 30, 30, 30, 30, 20 and 20 GMPCPP seeds, respectively. Data were pooled across 1–3 experiments. Error bars represent s.e.m.

Supplementary Figure 6 Controls and model for effects of tubulin depletion by nocodazole on nucleation in cells in Fig. 7.

(A) Plot of the microtubule growth rate against nocodazole concentration in two cell lines (U2OS, blue; LLCPK1, red). The microtubule growth rates were inferred from the velocity of EB comets. For increasing nocodazole concentrations, n = 25(2 independent experiments), 13(2), 13(2) and 10(2) (for LLCPK1 cells) and 20(2), 17(2), 12(2) and 12(2) (for U2OS cells) EB comets. Error bars represent the s.d. (B) Plot of the percentage of free soluble tubulin as a function of nocodazole concentration. Assuming a 1:1 stoichiometry and equilibrium conditions, we model the percentage of free tubulin as % Free Tubuli n = 100/(Keq−1[Nocodazole] + 1), where Keq is the equilibrium constant. The range of measured equilibrium constants is indicated. (C) Theoretical plot of the centrosomal nucleation rate (comets min−1 emerging from the centrosome) against predicted tubulin concentration in two cell lines (U2OS, blue; LLCPK1, red). The plot assumes an equilibrium constant for tubulin:nocodazole binding of 1 μM and a baseline soluble tubulin concentration of 10 μM. (D) Plot of the centrosomal nucleation rate against colchicine concentration in two cell lines (U2OS, blue; LLCPK1, red). For increasing colchicine concentrations, n = 12, 19, 18 and 19 U2OS cells, respectively and n = 29, 17, 20, 22 and 22 LLCPK1 cells, respectively. Data were pooled across 2 experiments. Error bars represent the s.d. (E) Kymographs of EB1-GFP end-tracking in vitro in control buffers (left), in the presence of 200 nM nocodazole (middle), or in the presence of 1 μM colchicine (right). (F) Images of Western blots performed against EB1 (left) and tubulin (right) on LLCPK1:EB1-GFP cells. The EB1 blot shows a band at 55 kDa in the LLCPK1:EB1-GFP cell line that is absent in the control LLCPK1 cells. We estimate EB1-GFP overexpression at 30% relative to the endogenous level based on the intensity of this band measured from two independent blots.

Supplementary Figure 7 Uncropped SDS-PAGE gels.

(A) Uncropped scanned image file of the TPX2 purification SDS-PAGE gel shown in Fig. 5a. (B) Uncropped scanned image file of the TPX2 cosedimentation assay SDS-PAGE gel shown in Supplementary Fig. 5A. (C) Uncropped scanned image file of the SUMO-tagged TPX2 truncation construct purification SDS-PAGE gel shown in Supplementary Fig. 5C. (D) Uncropped scanned image file of the MAP purification gel shown in Supplementary Fig. 5D.

Supplementary information

Supplementary Information

Supplementary Information (PDF 758 kb)

Microtubule nucleation from GMPCPP seeds at 12 μM tubulin.

Epifluorescence images of GMPCPP microtubule seeds (magenta) combined with total-internal-reflection fluorescence images of elongating microtubules (green) were recorded at 10 s intervals for 15 min. Some GMPCPP seeds are observed to produce microtubules immediately while others produce microtubules after a time lag. Some GMPCPP seeds do not produce microtubules during the experiment. Video playback is 100× real-time (see time stamp). (AVI 12756 kb)

Hysteresis in microtubule elongation experiments.

Epifluorescence images of GMPCPP microtubule seeds (magenta) combined with total-internal-reflection fluorescence images of elongating microtubules (green) were recorded at 10 s intervals. At the start of the experiment, the GMPCPP seeds are exposed to 4 μM tubulin (indicated). After a short period, the solution is exchanged with 15 μM tubulin (indicated). At this concentration, the GMPCPP seeds produce microtubules readily. The solution is exchanged back to 4 μM tubulin (indicated). Microtubules continue to elongate until they undergo catastrophe, after which the GMPCPP seed is dormant. Video playback is 100× real-time (see time stamp). (AVI 2544 kb)

LLCPK1 cells expressing GFP-EB1.

Spinning-disk confocal images of LLCPK1 cells constitutively expressing GFP-EB1. Images were taken every 2 s. Playback is 20× real-time (see time stamp). (AVI 19090 kb)

LLCPK1 cells expressing GFP-EB1 in the presence of 40 nM nocodazole.

Spinning-disk confocal images of LLCPK1 cells constitutively expressing GFP-EB1 in the presence of 40 nM nocodazole. Fewer ‘comets’ emerge from the centrosome. Images were taken every 2 s. Playback is 20× real-time (see time stamp). (AVI 19090 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wieczorek, M., Bechstedt, S., Chaaban, S. et al. Microtubule-associated proteins control the kinetics of microtubule nucleation. Nat Cell Biol 17, 907–916 (2015). https://doi.org/10.1038/ncb3188

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ncb3188

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