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:

Phase separation of EB1 guides microtubule plus-end dynamics

An Author Correction to this article was published on 07 December 2023

This article has been updated

Abstract

In eukaryotes, end-binding (EB) proteins serve as a hub for orchestrating microtubule dynamics and are essential for cellular dynamics and organelle movements. EB proteins modulate structural transitions at growing microtubule ends by recognizing and promoting an intermediate state generated during GTP hydrolysis. However, the molecular mechanisms and physiochemical properties of the EB1 interaction network remain elusive. Here we show that EB1 formed molecular condensates through liquid–liquid phase separation (LLPS) to constitute the microtubule plus-end machinery. EB1 LLPS is driven by multivalent interactions among different segments, which are modulated by charged residues in the linker region. Phase-separated EB1 provided a compartment for enriching tubulin dimers and other plus-end tracking proteins. Real-time imaging of chromosome segregation in HeLa cells expressing LLPS-deficient EB1 mutants revealed the importance of EB1 LLPS dynamics in mitotic chromosome movements. These findings demonstrate that EB1 forms a distinct physical and biochemical membraneless-organelle via multivalent interactions that guide microtubule dynamics.

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

Access options

Buy this article

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

Fig. 1: EB1 undergoes phase separation in live cells and in vitro.
Fig. 2: Lysine/arginine residues in the IDR encode molecular grammar for EB1 phase separation.
Fig. 3: EB1 phase separation regulates microtubule dynamics.
Fig. 4: EB1 phase separation is required for microtubule plus-end tracking activity.
Fig. 5: EB1 LLPS guides microtubule plus-end tracking machinery function in cells.
Fig. 6: Phase separation of EB1 ensures accurate chromosome segregation during mitosis.

Similar content being viewed by others

Data availability

Previously published protein sequence data that were re-analysed here are available from UniProt under accession codes EB1_Homo sapiens (ID: Q15691), EB1_Bovin (ID Q3ZBD9) EB1_Mouse (ID: Q61166), EB1_Xenopus (ID: Q6P848), EB1_Chicken (ID: Q5ZLC7), EB1_Danre (ID: Q6NUY9), EB1_fly (ID: A1Z6P3) and Mal3_yeast (ID: Q10113). All nucleotide sequences used in the design or synthesis were from genome annotation databases under accession codes EB1_Homo sapiens (ID: ENSG00000101367), EB1_Bovin (ID: ENSBTAT00000044474.4) EB1_Mouse (ID: ENSMUSG00000027479), EB1_Xenopus (ID: ENSXETT00000055099.4), EB1_Chicken (ID: ENSGALG00000006657), EB1_Danre (ID: ENSDART00000004474.9), EB1_fly (ID: FBtr0300285) and Mal3_yeast (ID: SPAC18G6.15.1). Previously published structures are available from the Protein Data Bank under accession codes EB1 CH domain (PDB entry 1PA7 (ref. 8)) and the EB1 EBH domain (PDB entry 3GJO26). Source data are provided with this paper. All other data supporting the findings of this study are available from the corresponding author on reasonable request.

Code availability

The custom code for coarse-grained computational modelling and statistical analysis along with its output dataset are available at http://zenodo.org with the identifier https://doi.org/10.5281/zenodo.7113746. RPA and shuffling data analysis is deposited at https://github.com/Carolinge/EB1_rpashuffle. Source data are provided with this paper.

Change history

References

  1. Brouhard, G. J. & Rice, L. M. Microtubule dynamics: an interplay of biochemistry and mechanics. Nat. Rev. Mol. Cell Biol. 19, 451–463 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  3. Akhmanova, A. & Steinmetz, M. O. Tracking the ends: a dynamic protein network controls the fate of microtubule tips. Nat. Rev. Mol. Cell Biol. 9, 309–322 (2008).

    Article  CAS  PubMed  Google Scholar 

  4. Dragestein, K. A. et al. Dynamic behavior of GFP-CLIP-170 reveals fast protein turnover on microtubule plus ends. J. Cell Biol. 180, 729–737 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Seetapun, D., Castle, B. T., McIntyre, A. J., Tran, P. T. & Odde, D. J. Estimating the microtubule GTP cap size in vivo. Curr. Biol. 22, 1681–1687 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Akhmanova, A. & Steinmetz, M. O. Control of microtubule organization and dynamics: two ends in the limelight. Nat. Rev. Mol. Cell Biol. 16, 711–726 (2015).

    Article  CAS  PubMed  Google Scholar 

  7. Bu, W. & Su, L. K. Characterization of functional domains of human EB1 family proteins. J. Biol. Chem. 278, 49721–49731 (2003).

    Article  CAS  PubMed  Google Scholar 

  8. Hayashi, I. & Ikura, M. Crystal structure of the amino-terminal microtubule-binding domain of end-binding protein 1 (EB1). J. Biol. Chem. 278, 36430–36434 (2003).

    Article  CAS  PubMed  Google Scholar 

  9. Jiang, K. et al. TIP150 interacts with and targets MCAK at the microtubule plus ends. EMBO Rep. 10, 857–865 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Lee, T., Langford, K. J., Askham, J. M., Bruning-Richardson, A. & Morrison, E. E. MCAK associates with EB1. Oncogene 27, 2494–2500 (2008).

    Article  CAS  PubMed  Google Scholar 

  11. 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 

  12. Perez, F., Diamantopoulos, G. S., Stalder, R. & Kreis, T. E. CLIP-170 highlights growing microtubule ends in vivo. Cell 96, 517–527 (1999).

    Article  CAS  PubMed  Google Scholar 

  13. Weisbrich, A. et al. Structure-function relationship of CAP-Gly domains. Nat. Struct. Mol. Biol. 14, 959–967 (2007).

    Article  CAS  PubMed  Google Scholar 

  14. Zhang, R., Alushin, G. M., Brown, A. & Nogales, E. Mechanistic Origin of Microtubule Dynamic Instability and Its Modulation by EB Proteins. Cell 162, 849–859 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Xia, P. et al. EB1 acetylation by P300/CBP-associated factor (PCAF) ensures accurate kinetochore-microtubule interactions in mitosis. Proc. Natl Acad. Sci. USA 109, 16564–16569 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Xia, P. et al. Superresolution imaging reveals structural features of EB1 in microtubule plus-end tracking. Mol. Biol. Cell 25, 4166–4173 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  17. Guo, Y. et al. Visualizing intracellular organelle and cytoskeletal interactions at nanoscale resolution on millisecond timescales. Cell 175, 1430–1442 e1417 (2018).

    Article  CAS  PubMed  Google Scholar 

  18. Liu, X. et al. Phase separation drives decision making in cell division. J. Biol. Chem. 295, 13419–13431 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Ong, J. Y. & Torres, J. Z. Phase separation in cell division. Mol. Cell 80, 9–20 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Musacchio, A. On the role of phase separation in the biogenesis of membraneless compartments. EMBO J. 41, e109952 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Trivedi, P. et al. The inner centromere is a biomolecular condensate scaffolded by the chromosomal passenger complex. Nat. Cell Biol. 21, 1127–1137 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Maan, R. et al. Multivalent interactions facilitate motor-dependent protein accumulation at growing microtubule plus ends. Nat. Cell Biol. https://doi.org/10.1038/s41556-022-01037-0 (2022).

  23. Shin, Y. et al. Spatiotemporal control of intracellular phase transitions using light-activated optoDroplets. Cell 168, 159–171 e114 (2017).

    Article  CAS  PubMed  Google Scholar 

  24. Tsang, B., Pritisanac, I., Scherer, S. W., Moses, A. M. & Forman-Kay, J. D. Phase separation as a missing mechanism for interpretation of disease mutations. Cell 183, 1742–1756 (2020).

    Article  CAS  PubMed  Google Scholar 

  25. Zeng, M. et al. Phase transition in postsynaptic densities underlies formation of synaptic complexes and synaptic plasticity. Cell 166, 1163–1175 e1112 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Honnappa, S. et al. An EB1-binding motif acts as a microtubule tip localization signal. Cell 138, 366–376 (2009).

    Article  CAS  PubMed  Google Scholar 

  27. Dignon, G. L., Zheng, W., Best, R. B., Kim, Y. C. & Mittal, J. Relation between single-molecule properties and phase behavior of intrinsically disordered proteins. Proc. Natl Acad. Sci. USA 115, 9929–9934 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Leicher, R. et al. Single-stranded nucleic acid binding and coacervation by linker histone H1. Nat. Struct. Mol. Biol. 29, 463–471 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Wang, J. et al. A molecular grammar governing the driving forces for phase separation of prion-like RNA binding proteins. Cell 174, 688–699 e616 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Kroschwald, S. et al. Different material states of Pub1 condensates define distinct modes of stress adaptation and recovery. Cell Rep. 23, 3327–3339 (2018).

    Article  CAS  PubMed  Google Scholar 

  31. Wu, Q., Schapira, M., Arrowsmith, C. H. & Barsyte-Lovejoy, D. Protein arginine methylation: from enigmatic functions to therapeutic targeting. Nat. Rev. Drug Discov. 20, 509–530 (2021).

    Article  CAS  PubMed  Google Scholar 

  32. Leslie, M. Separation anxiety. Science 371, 336–338 (2021).

    Article  CAS  PubMed  Google Scholar 

  33. Jiang, K. et al. A proteome-wide screen for mammalian SxIP motif-containing microtubule plus-end tracking proteins. Curr. Biol. 22, 1800–1807 (2012).

    Article  CAS  PubMed  Google Scholar 

  34. Lin, Y. H., Forman-Kay, J. D. & Chan, H. S. Sequence-specific polyampholyte phase separation in membraneless organelles. Phys. Rev. Lett. 117, 178101 (2016).

    Article  PubMed  Google Scholar 

  35. 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 

  36. Maurer, S. P. et al. EB1 accelerates two conformational transitions important for microtubule maturation and dynamics. Curr. Biol. 24, 372–384 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Vaughan, K. T. TIP maker and TIP marker; EB1 as a master controller of microtubule plus ends. J. Cell Biol. 171, 197–200 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Komarova, Y. et al. Mammalian end binding proteins control persistent microtubule growth. J. Cell Biol. 184, 691–706 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Kim, H. J. et al. Mutations in prion-like domains in hnRNPA2B1 and hnRNPA1 cause multisystem proteinopathy and ALS. Nature 495, 467–473 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Molliex, A. et al. Phase separation by low complexity domains promotes stress granule assembly and drives pathological fibrillization. Cell 163, 123–133 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Banani, S. F. et al. Compositional control of phase-separated cellular bodies. Cell 166, 651–663 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. McKinley, K. L. & Cheeseman, I. M. Large-scale analysis of CRISPR/Cas9 cell-cycle knockouts reveals the diversity of p53-dependent responses to cell-cycle defects. Dev. Cell 40, 405–420 e402 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Mo, F. et al. Acetylation of Aurora B by TIP60 ensures accurate chromosomal segregation. Nat. Chem. Biol. 12, 226–232 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Ward, T. et al. Regulation of a dynamic interaction between two microtubule-binding proteins, EB1 and TIP150, by the mitotic p300/CBP-associated factor (PCAF) orchestrates kinetochore microtubule plasticity and chromosome stability during mitosis. J. Biol. Chem. 288, 15771–15785 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Meier, S. M. et al. Multivalency ensures persistence of a +TIP body at specialized microtubule ends. Nat. Cell Biol. https://doi.org/10.1038/s41556-022-01035-2 (2022).

  46. Song, X. et al. Dynamic crotonylation of EB1 by TIP60 ensures accurate spindle positioning in mitosis. Nat. Chem. Biol. 17, 1314–1323 (2021).

    Article  CAS  PubMed  Google Scholar 

  47. Yao, X., Abrieu, A., Zheng, Y., Sullivan, K. F. & Cleveland, D. W. CENP-E forms a link between attachment of spindle microtubules to kinetochores and the mitotic checkpoint. Nat. Cell Biol. 2, 484–491 (2000).

    Article  CAS  PubMed  Google Scholar 

  48. Li, P. et al. Phase transitions in the assembly of multivalent signalling proteins. Nature 483, 336–340 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Shcherbakov, A. A., Spreacker, P. J., Dregni, A. J., Henzler-Wildman, K. A. & Hong, M. High-pH structure of EmrE reveals the mechanism of proton-coupled substrate transport. Nat. Commun. 13, 991 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Huang, Y. et al. BubR1 phosphorylates CENP-E as a switch enabling the transition from lateral association to end-on capture of spindle microtubules. Cell Res. 29, 562–578 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Polverino, F. et al. The Aurora-A/TPX2 axis directs spindle orientation in adherent human cells by regulating NuMA and microtubule stability. Curr. Biol. 31, 658–667 e655 (2021).

    Article  CAS  PubMed  Google Scholar 

  52. Blas, F. J., MacDowell, L. G., de Miguel, E. & Jackson, G. Vapor–liquid interfacial properties of fully flexible Lennard–Jones chains. J. Chem. Phys. 129, 144703 (2008).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We are grateful to Mingjie Zhang, Peace Cheng, Yeguang Chen and Yi Luo for critical reading, and to Iain M. Cheeseman, Li Yu and Clifford P. Brangwynne for reagents. This work was supported by MOST-NSFC grants (2017YFA0503600, 32090040, 91854203, 31621002, 21922706, 92153302, 92254302, 2022YFA1303100, 2022YFA0806800 and 91853115 to X.L.; 92059102 and 92253305 to X.S.; 2019YFA0508403 and 31971128 to S.X.; 92253301 to C.X.; 2022YFA1302700 to Z.W.); the Ministry of Education (IRT_17R102, 20113402130010 and YD2070006001) to X.L.; the Strategic Priority Research Program of the Chinese Academy of Sciences (XDB19040000 to X.L.; XDB37040202 to S.X.); the Fundamental Research Funds for the Central Universities (WK2070000194) to X.L.; Anhui Provincial Natural Science Foundation Grant (1908085MC64) and China Postdoctoral Science Foundation Grant (2019M662184) to X.S., and NIH grants (CA146133 and DK115812). The funders had no role in study design, data collection and analysis, decision to publish or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

X.Y. and X.L. conceived the project. X.S., T.Y. and M. Ding designed and performed most of the biochemical experiments. X.S. performed OptoDroplet assay, and F.Y. performed live-cell imaging and IF analyses. X.S. and D.L. performed GI-SIM and lattice-SIM imaging. Y.W., Z.W. and K.J. performed in vitro TIRF assays. Linge L. and Z.H. performed molecular modelling. P.X. and S.X. performed NMR experiments. T.Y. and C.X. performed ITC assays. C.C. performed the surface charge analysis and prepared structure-modelling images. S.L. and M. Dai engineered the EB1-GFP knock-in HeLa cell line. X.S., Y.W. and F.Y. performed data analyses. P.L. contributed reagents. X.S., X.L. and X.Y. wrote the manuscript. K.Y., C.F., J.Z., Y.S., Lin L. and D.L.H. edited the manuscript. All authors have commented on and approved the manuscript.

Corresponding authors

Correspondence to Xing Liu or Xuebiao Yao.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Cell Biology thanks Carsten Janke, Claire Walczak and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 EB1 comets exhibit liquid-like property.

(a) GI-SIM image of a typical GFP-EB1 comet fusion at the microtubule plus-ends. COS-7 cells were transfected with EB1-GFP and mCherry-tubulin. The boxed area is magnified in the right panels, which shows time-lapse imaging of EB1 comets over 4 s. White arrows highlight the fusion process of two adjacent comets. Scale bar, 10 μm in original image and 5 μm in magnified montages. (b) Time-lapse analysis of EB1-GFP in HeLa cells by Lattice-SIM. A representative cell with EB1 comet fusion (arrowheads) at different z-plane was shown. Scale bar, 10 μm. (c) Schematic representation of engineering design of the endogenous EB1 gene with an mEGFP tag via CRISPR-Cas9 mediated gene editing to generate EB1-GFP knock-in (KI) in HeLa cells. The mEGFP were inserted before the TAA stop codon of EB1. (d) The EB1-GFP KI HeLa cells were verified by Western blotting using an anti-EB1 antibody. (e) Time-lapse analysis of EB1-GFP KI HeLa cells by Lattice-SIM. The arrows indicate EB1 comets before and after fusion. Scale bar: in original images, 10 μm; in zoom-in images, 5 μm. (f) Schematic of the OptoEB1 construct, which contains mCherry and Cry2 (blue light-induced oligomerization). (g) Representative snapshots of mCherry-Cry2 (mCh-Cry2) in living COS-7 cells upon blue light activation. Note that Cry2 alone does not exhibit comet-like structure after blue-light activation. Scale bar, 10 μm.

Source data

Extended Data Fig. 2 Phase separation property of EB1 is evolutionarily conserved in eukaryotic cells.

(a) Sequence alignment of EB1 from yeast to human. The residues with positive charge were highlighted. (b) Schematic representation of the domain structures of human EB1, fission yeast Mal3 and budding yeast Bim1. (c) In vitro phase separation assay. Representative fluorescence images of Mal3-GFP phase separation in BRB80 buffer with various concentrations of KCl. Scale bar, 5 µm. (d) Representative SDS-PAGE gel of EB1 solubility in solution. Samples of EB1-GFP protein at various concentration were centrifuged as described under “Methods”. Equal volumes of supernatant (S) and pellet (P) fractions were resolved by electrophoresis and resolved proteins were visualized by CBB staining. (e) Determination of a saturation concentration (csat) of EB1-GFP. The absorbance at 340 nm (condensed protein has a distinctly different absorbance profile, leading to a pronounced uptick in A340 absorbance) was measured. The tendency of soluble and condensed concentration were labeled by red line, and the intersection was measured as csat. Data represent mean ± s.e.m. from three independent experiments. (f) In vitro phase separation assay. Representative fluorescence and DIC images of EB1-GFP (20 μM) phase separation with or without 5% 1,6-Hexanediol (Hex) in BRB80 buffer with 150 mM KCl. Scale bar, 5 µm. (g) Schematic illustration of an ultra-coarse-grained modeling for CH domain, EBH domain, the IDR region and C-terminal region. (h) Representative image of EB1 phase separation based on the ultra-coarse-grained modeling.

Source data

Extended Data Fig. 3 Characterization of multivalent interactions among EB1 regions by NMR.

(a) Combined chemical shifts perturbation (CSP) of IDR in the presence of EBH (IDR vs IDR + EBH). The dash lines indicated the level of standard deviation values. Proteins boxed in red were labeled by 15N. Gray bars indicated overlapping or unassigned peaks, while the asterisks labeled the proline residues which had no signal in NMR assay. (b) Combined chemical shifts perturbation of IDR + EBH in the presence of CTT (IDR + EBH vs IDR + EBH + CTT. *, proline residues. (c) Peak intensity change of IDR + EBH + CTT (15N labeling) upon CH domain titration. The error bars were calculated based on every peak’s signal-to-noise ratios as described in Methods. *, proline residues. (d) Representative spectra of indicated regions were overlaid to show the peak shifts. The spectra of IDR, IDR + EBH, and IDR + EBH + CTT were shown in red, blue, and green, respectively. The peak shifts of residues showed the interaction between indicated regions. (e) Summary of the interactions among different EB1 regions.

Source data

Extended Data Fig. 4 Lysine/arginine residues on IDR are essential for EB1 phase separation.

(a) Characterization of recombinant purified proteins used in in vitro phase separation assay. Recombinant EB1 proteins were subjected to electrophoresis followed by CBB staining. (b) Diagram of EB1 wild type and IDR chimeras. Their phase separation capacities at 10 μM were also annotated. “−” no phase separation events were observed; “+” phase separation events were observed. (c) Representative images of EB1 chimeras in COS-7 cells (Upper) and corresponding phase separation test in vitro (Lower). Scale bar, 10 μm. (d) Sequence alignment of EB1-IDR and Pub1-IDR. The residues with positive charge were highlighted. (e) Summary of in vitro phase separation behavior of EB1 RDI and KR6A mutants. EB1 mutants were purified and assessed for in vitro phase separation assay. The saturation concentration was measured by turbidity assay at indicated salt conditions. The red circles indicate phase separation (above saturation concentration), and the blue circles indicate no phase separation (below saturation concentration). (f) EB1 binding with purified +TIPs in vitro. Purified GST- or MBP-tagged +TIPs were used as affinity matrices to absorb His-EB1 wild type or KR6Q mutant as in Fig. 2k. The bound His-EB1 protein was assessed by Western blotting using an anti-His antibody.

Source data

Extended Data Fig. 5 Binding affinity between EB1 and its binding partners was not changed by KR6Q mutation.

ITC binding curves for interaction of EB1 and its binding partners upon KR6Q mutation. (a) MCAK peptide, aa 85-118; (b) TIP150 peptide, aa 823-856; (c) CLASP1 peptide, aa 713-747; (d) CLASP2 peptide, aa 490-523; (e) p150Glued CAP-GLY domain (aa 1-105). For ITC assay, the peptides and p150Glued CAP-GLY domain were used at 1 mM in syringe, while the purified EB1 wild type and KR6Q mutant were used at 50 μM in cell. Data are representative from three independent replicates. All peptide sequences were listed in Supplementary Table 1.

Extended Data Fig. 6 Characterization of positive charges and their positions in IDR for EB1 phase separation.

(a) Schematic diagram of three classes of EB1 mutants: 1) positive charges in IDR were mutated (KR1Q-KR6Q); 2) the positions of charges in IDR were rearranged in cluster as 1-10 K/R; 3) a scramble sequence of IDR (scr9). The positive charges and mutations in IDR were annotated. All the sequences of EB1 mutants were listed in Supplementary Table 2. (b) Ultra-coarse-grained computational modeling to elucidate the phase separation behaviors of the three classes of EB1 mutants as in (a). (c) The random phase approximation (RPA) theory was applied to estimate the IDR electrostatic effects on EB1 phase separation. (d) The in vitro phase separation assay of EB1 mutants (10 μM) in the presence of 150 mM KCl. Scale bar, 5 µm. (e) The phase diagram of EB1 mutants generated by in vitro phase separation assay. The saturation concentration was measured by turbidity assay at indicated salt and protein concentrations. The red circles indicate phase separation (above saturation concentration), and the blue circles indicate no phase separation (below saturation concentration). Data are representative from three independent replicates.

Extended Data Fig. 7 Phase separation diagram of EB1 from different species in vitro.

Summary of in vitro phase separation behavior of wild-type and lysine/arginine mutated EB1 in response to changes in salt concentration. Proteins of both wild-type and mutated EB1 from different species were purified and assessed for in vitro phase separation propensity as described in Methods. The saturation concentration was measured by turbidity assay at indicated salt and protein concentrations. The red circles indicate phase separation (above saturation concentration), and the blue circles indicate no phase separation (below saturation concentration). Data are representative from three independent replicates.

Extended Data Fig. 8 Characterization of EB1 phase separation in the presence of microtubule and +TIPs in vitro.

(a) Characterization of EB1 phase separation in microtubule plus-end tracking at different KCl concentrations using in vitro TIRFM assay. The concentration of EB1-GFP was 50 nM in the presence of various concentrations of KCl. Scale bar, 5 μm. (b-c) Quantitative analyses of fluorescence intensity of EB1 at microtubule plus-ends (b) and microtubule growth rate (c) in (a). For each group, n = 30 microtubules pooled from three independent experiments. Data represent mean ± s.e.m. Ordinary one-way ANOVA followed by Tukey’s post hoc test was used to determine statistical significance. ****p < 0.0001; ns (not significant) indicates p > 0.05. (d) Electron micrograph of EB1-hnRNPA1 chimeras (2 μM) from in vitro phase separation assay as carried out in Fig. 4a. Scale bar, 500 nm. (e) Representative images of co-phase separation of FITC-labeled His-EB1 (EB1-FITC) and other +TIPs co-condensates in the presence of 150 mM KCl. Scale bar, 5 μm. (f) Partition coefficients (PCs) for EB1 and other +TIPs from experiment in (e). Data represent mean ± s.e.m. from three independent experiments. Dotted line, PC = 1. (g) Quantitative analyses of droplet areas in (e). For EB1-FITC, n = 30; EB1-FITC + TIP150, n = 50; EB1-FITC + MCAK, n = 30; EB1-FITC + p150Glued, n = 30; EB1-FITC + CLIP170, n = 30; for +TIPs only, no droplet was observed in ten fields of view under microscope; n represents the number of EB1 droplets pooled from three independent experiments. Data represent mean ± s.e.m. Ordinary one-way ANOVA followed by Tukey’s post hoc test was used to determine statistical significance. ****p < 0.0001; ns (not significant) indicates p > 0.05.

Source data

Extended Data Fig. 9 Phase separation is a function of EB1 in microtubule plus-end tracking.

(a-b) Representative immunofluorescence images of FLAG-EB1 (wild type and KR6Q) and GFP-tagged +TIPs proteins (MCAK and p150Glued) co-expressing in endogenous EB1-depleted HeLa cells. DNA was stained with DAPI and microtubule was stained with an anti-tubulin antibody. Scale bar, 10 μm.

Extended Data Fig. 10 Dynamic phase separation of EB1 guides chromosome movements.

(a) Characterization of expression levels of GFP-tagged EB1 wild type, KR6Q mutant and EB1-hnRNPA1 chimeras in endogenous EB1-depleted HeLa cells were analyzed by Western blot with an anti-EB1 antibody. Doxycycline (Dox) was added to induce EB1 knockout. (b) Representative mitotic phenotypes in endogenous EB1-depleted HeLa cells expressing EB1-hnRNPA1 chimeras shown by time-lapse microscopy. Chromosomes were visualized with mCherry-H2B. Arrows, spindle poles; arrowheads, lagging chromosomes. Scale bar, 10 μm. (c) Quantification of mitotic phenotypes of live HeLa cells in (b). For each group, n = 30 cells pooled from three independent experiments. Data represent mean ± s.e.m. Statistical significance was assessed by two-way ANOVA. ****p < 0.0001; ***p = 0.001; **p = 0.0041; *p = 0.0107; ns (not significant) indicates p > 0.05.

Source data

Supplementary information

Reporting Summary

Supplementary Tables

Sequence of peptides for ITC assay and sequence of EB1 mutants.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 2

Unprocessed WBs and/or gels.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 1

Unprocessed WBs and/or gels.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 2

Unprocessed WBs and/or gels.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Unprocessed WBs and/or gels.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Source Data Extended Data Fig. 10

Unprocessed WBs and/or gels.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Song, X., Yang, F., Yang, T. et al. Phase separation of EB1 guides microtubule plus-end dynamics. Nat Cell Biol 25, 79–91 (2023). https://doi.org/10.1038/s41556-022-01033-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-022-01033-4

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