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:

Extreme dynamics in a biomolecular condensate

Abstract

Proteins and nucleic acids can phase-separate in the cell to form concentrated biomolecular condensates1,2,3,4. The functions of condensates span many length scales: they modulate interactions and chemical reactions at the molecular scale5, organize biochemical processes at the mesoscale6 and compartmentalize cells4. Understanding the underlying mechanisms of these processes will require detailed knowledge of the rich dynamics across these scales7. The mesoscopic dynamics of biomolecular condensates have been extensively characterized8, but their behaviour at the molecular scale has remained more elusive. Here, as an example of biomolecular phase separation, we study complex coacervates of two highly and oppositely charged disordered human proteins9. Their dense phase is 1,000 times more concentrated than the dilute phase, and the resulting percolated interaction network10 leads to a bulk viscosity 300 times greater than that of water. However, single-molecule spectroscopy optimized for measurements within individual droplets reveals that at the molecular scale, the disordered proteins remain exceedingly dynamic, with their chain configurations interconverting on submicrosecond timescales. Massive all-atom molecular dynamics simulations reproduce the experimental observations and explain this apparent discrepancy: the underlying interactions between individual charged side chains are short-lived and exchange on a pico- to nanosecond timescale. Our results indicate that, despite the high macroscopic viscosity of phase-separated systems, local biomolecular rearrangements required for efficient reactions at the molecular scale can remain rapid.

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: Mesoscopic and microscopic properties of ProTα–H1 droplets.
Fig. 2: Single-molecule spectroscopy in the dilute and dense phases.
Fig. 3: Large-scale MD simulations of ProTα–H1 phase separation.
Fig. 4: The complex hierarchy of length- and timescales in phase-separated droplets.

Similar content being viewed by others

Data availability

The simulation trajectories of the condensates have been deposited at Zenodo (https://doi.org/10.5281/zenodo.7963359).  Source data are provided with this paper.

Code availability

Fretica, a custom add-on package for Mathematica v.12.3 (Wolfram Research) was used for the analysis of single-molecule fluorescence data and is available at https://github.com/SchulerLab. The code used to calculate the lifetime of residue–residue contacts is available at https://doi.org/10.5281/zenodo.7967716.

References

  1. Gibson, B. A. et al. Organization of chromatin by intrinsic and regulated phase separation. Cell 179, 470–484 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Gibbs, E. B. & Kriwacki, R. W. Linker histones as liquid-like glue for chromatin. Proc. Natl Acad. Sci. USA 115, 11868–11870 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  3. Shin, Y. & Brangwynne, C. P. Liquid phase condensation in cell physiology and disease. Science 357, eaaf4382 (2017).

    Article  PubMed  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  PubMed  PubMed Central  Google Scholar 

  5. Nakashima, K. K., Vibhute, M. A. & Spruijt, E. Biomolecular chemistry in liquid phase separated compartments. Front. Mol. Biosci. 6, 21 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Snead, W. T. & Gladfelter, A. S. The control centers of biomolecular phase separation: how membrane surfaces, PTMs, and active processes regulate condensation. Mol. Cell 76, 295–305 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Lyon, A. S., Peeples, W. B. & Rosen, M. K. A framework for understanding the functions of biomolecular condensates across scales. Nat. Rev. Mol. Cell Biol. 22, 215–235 (2021).

    Article  CAS  PubMed  Google Scholar 

  8. Alshareedah, I., Kaur, T. & Banerjee, P. R. Methods for characterizing the material properties of biomolecular condensates. Methods Enzymol. 646, 143–183 (2021).

    Article  CAS  PubMed  Google Scholar 

  9. Borgia, A. et al. Extreme disorder in an ultrahigh-affinity protein complex. Nature 555, 61–66 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  10. Farag, M. et al. Condensates formed by prion-like low-complexity domains have small-world network structures and interfaces defined by expanded conformations. Nat. Commun. 13, 7722 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  11. Vendruscolo, M. & Fuxreiter, M. Protein condensation diseases: therapeutic opportunities. Nat. Commun. 13, 5550 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  12. Toretsky, J. A. & Wright, P. E. Assemblages: functional units formed by cellular phase separation. J. Cell Biol. 206, 579–588 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Brangwynne, C. P., Tompa, P. & Pappu, R. V. Polymer physics of intracellular phase transitions. Nat. Phys. 11, 899–904 (2015).

    Article  CAS  Google Scholar 

  14. Ruff, K. M., Pappu, R. V. & Holehouse, A. S. Conformational preferences and phase behavior of intrinsically disordered low complexity sequences: insights from multiscale simulations. Curr. Opin. Struct. Biol. 56, 1–10 (2019).

    Article  CAS  PubMed  Google Scholar 

  15. Dignon, G. L., Best, R. B. & Mittal, J. Biomolecular phase separation: from molecular driving forces to macroscopic properties. Annu. Rev. Phys. Chem. 71, 53–75 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  16. Murthy, A. C. & Fawzi, N. L. The (un)structural biology of biomolecular liquid-liquid phase separation using NMR spectroscopy. J. Biol. Chem. 295, 2375–2384 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Ahmed, R. & Forman-Kay, J. D. NMR insights into dynamic, multivalent interactions of intrinsically disordered regions: from discrete complexes to condensates. Essays Biochem. 66, 863–873 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Alberti, S., Gladfelter, A. & Mittag, T. Considerations and challenges in studying liquid-liquid phase separation and biomolecular condensates. Cell 176, 419–434 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Wei, M. T. et al. Phase behaviour of disordered proteins underlying low density and high permeability of liquid organelles. Nat. Chem. 9, 1118–1125 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Jawerth, L. et al. Protein condensates as aging Maxwell fluids. Science 370, 1317–1323 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  21. Nasir, I., Onuchic, P. L., Labra, S. R. & Deniz, A. A. Single-molecule fluorescence studies of intrinsically disordered proteins and liquid phase separation. Biochim. Biophys. Acta Proteins Proteom. 1867, 980–987 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Mazal, H. & Haran, G. Single-molecule FRET methods to study the dynamics of proteins at work. Curr. Opin. Biomed. Eng. 12, 8–17 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  23. Lerner, E. et al. FRET-based dynamic structural biology: challenges, perspectives and an appeal for open-science practices. eLife 10, e60416 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Schuler, B., Soranno, A., Hofmann, H. & Nettels, D. Single-molecule FRET spectroscopy and the polymer physics of unfolded and intrinsically disordered proteins. Annu. Rev. Biophys. 45, 207–231 (2016).

    Article  CAS  PubMed  Google Scholar 

  25. Schuler, B. et al. Binding without folding—the biomolecular function of disordered polyelectrolyte complexes. Curr. Opin. Struct. Biol. 60, 66–76 (2020).

    Article  CAS  PubMed  Google Scholar 

  26. Heidarsson, P. O. et al. Release of linker histone from the nucleosome driven by polyelectrolyte competition with a disordered protein. Nat. Chem. 14, 224–231 (2022).

    Article  CAS  PubMed  Google Scholar 

  27. Shakya, A., Park, S., Rana, N. & King, J. T. Liquid-liquid phase separation of histone proteins in cells: role in chromatin organization. Biophys. J. 118, 753–764 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  28. Sottini, A. et al. Polyelectrolyte interactions enable rapid association and dissociation in high-affinity disordered protein complexes. Nat. Commun. 11, 5736 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  29. Srivastava, S. & Tirrell, M. V. Polyelectrolyte complexation. Adv. Chem. Phys. 161, 499–544 (2016).

    CAS  Google Scholar 

  30. Rumyantsev, A. M., Jackson, N. E. & de Pablo, J. J. Polyelectrolyte complex coacervates: recent developments and new frontiers. Annu. Rev. Condens. Matter Phys. 12, 155–176 (2021).

    Article  ADS  Google Scholar 

  31. Fisher, R. S. & Elbaum-Garfinkle, S. Tunable multiphase dynamics of arginine and lysine liquid condensates. Nat. Commun. 11, 4628 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  32. Martin, E. W. et al. Valence and patterning of aromatic residues determine the phase behavior of prion-like domains. Science 367, 694–699 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  33. Record, M. T. Jr., Anderson, C. F. & Lohman, T. M. Thermodynamic analysis of ion effects on the binding and conformational equilibria of proteins and nucleic acids: the roles of ion association or release, screening, and ion effects on water activity. Q. Rev. Biophys. 11, 103–178 (1978).

    Article  CAS  PubMed  Google Scholar 

  34. Wang, H., Kelley, F. M., Milovanovic, D., Schuster, B. S. & Shi, Z. Surface tension and viscosity of protein condensates quantified by micropipette aspiration. Biophys. Rep. 1, 100011 (2021).

    CAS  Google Scholar 

  35. Tuinier, R., Dhont, J. K. G. & Fan, T. H. How depletion affects sphere motion through solutions containing macromolecules. Europhys. Lett. 75, 929–935 (2006).

  36. Rubinstein, M. & Colby, R. H. Polymer Physics (Oxford Univ. Press, 2003).

  37. Muthukumar, M. Physics of Charged Macromolecules: Synthetic and Biological Systems (Cambridge Univ. Press, 2023).

  38. Muthukumar, M. Dynamics of polyelectrolyte solutions. J. Chem. Phys. 107, 2619–2635 (1997).

    Article  ADS  CAS  Google Scholar 

  39. Aznauryan, M. et al. Comprehensive structural and dynamical view of an unfolded protein from the combination of single-molecule FRET, NMR, and SAXS. Proc. Natl Acad. Sci. USA 113, E5389–5398 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Zheng, W. et al. Inferring properties of disordered chains from FRET transfer efficiencies. J. Chem. Phys. 148, 123329 (2018).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  41. Brady, J. P. et al. Structural and hydrodynamic properties of an intrinsically disordered region of a germ cell-specific protein on phase separation. Proc. Natl Acad. Sci. USA 114, E8194–E8203 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Soranno, A. et al. Quantifying internal friction in unfolded and intrinsically disordered proteins with single-molecule spectroscopy. Proc. Natl Acad. Sci. USA 109, 17800–17806 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  43. Zheng, W. et al. Molecular details of protein condensates probed by microsecond long atomistic simulations. J. Phys. Chem. B 124, 11671–11679 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Best, R. B., Zheng, W. & Mittal, J. Balanced protein-water interactions improve properties of disordered proteins and non-specific protein association. J. Chem. Theory Comput. 10, 5113–5124 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Abascal, J. L. F. & Vega, C. A general purpose model for the condensed phases of water: TIP4P/2005. J. Chem. Phys. 123, 234505 (2005).

    Article  ADS  CAS  PubMed  Google Scholar 

  46. Shea, J. E., Best, R. B. & Mittal, J. Physics-based computational and theoretical approaches to intrinsically disordered proteins. Curr. Opin. Struct. Biol. 67, 219–225 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Nüesch, M. F. et al. Single-molecule detection of ultrafast biomolecular dynamics with nanophotonics. J. Am. Chem. Soc. 144, 52–56 (2022).

    Article  PubMed  Google Scholar 

  48. Guenza, M. Cooperative dynamics in unentangled polymer fluids. Phys. Rev. Lett. 88, 025901 (2002).

    Article  ADS  CAS  PubMed  Google Scholar 

  49. Reinkemeier, C. D. & Lemke, E. A. Synthetic biomolecular condensates to engineer eukaryotic cells. Curr. Opin. Chem. Biol. 64, 174–181 (2021).

    Article  CAS  PubMed  Google Scholar 

  50. Wen, J. et al. Conformational expansion of tau in condensates promotes irreversible aggregation. J. Am. Chem. Soc. 143, 13056–13064 (2021).

    Article  CAS  PubMed  Google Scholar 

  51. Bottaro, S. & Lindorff-Larsen, K. Biophysical experiments and biomolecular simulations: a perfect match? Science 361, 355–360 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  52. Koenig, I. et al. Single-molecule spectroscopy of protein conformational dynamics in live eukaryotic cells. Nat. Methods 12, 773–779 (2015).

    Article  Google Scholar 

  53. Perry, S. L. Phase separation: bridging polymer physics and biology. Curr. Opin. Colloid Interface Sci. 39, 86–97 (2019).

    Article  CAS  Google Scholar 

  54. Lin, Y. et al. Narrow equilibrium window for complex coacervation of tau and RNA under cellular conditions. eLife 8, e42571 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  55. Yeh, I. C. & Hummer, G. System-size dependence of diffusion coefficients and viscosities from molecular dynamics simulations with periodic boundary conditions. J. Phys. Chem. B 108, 15873–15879 (2004).

    Article  CAS  Google Scholar 

  56. Chowdhury, A. et al. Mechanism-dependent modulation of ultrafast interfacial water dynamics in intrinsically disordered protein complexes. Angew. Chem. Int. Edn 58, 4720–4724 (2019).

    Article  CAS  Google Scholar 

  57. Müller, B. K., Zaychikov, E., Bräuchle, C. & Lamb, D. C. Pulsed interleaved excitation. Biophys. J. 89, 3508–3522 (2005).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Schuler, B. Application of single molecule Förster resonance energy transfer to protein folding. Methods Mol. Biol. 350, 115–138 (2007).

    CAS  PubMed  Google Scholar 

  59. Hellenkamp, B. et al. Precision and accuracy of single-molecule FRET measurements-a multi-laboratory benchmark study. Nat. Methods 15, 669–676 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Van Der Meer, B. W., Coker, G. III & Chen, S. Y. S. Resonance Energy Transfer: Theory and Data (VCH Publishers, 1994).

  61. Klose, D. et al. Resolving distance variations by single-molecule FRET and EPR spectroscopy using rotamer libraries. Biophys. J. 120, 4842–4858 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  62. Zhao, H., Brown, P. H. & Schuck, P. On the distribution of protein refractive index increments. Biophys. J. 100, 2309–2317 (2011).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  63. Barer, R. & Tkaczyk, S. Refractive index of concentrated protein solutions. Nature 173, 821–822 (1954).

    Article  ADS  CAS  PubMed  Google Scholar 

  64. Haritos, A. A., Tsolas, O. & Horecker, B. L. Distribution of prothymosin alpha in rat tissues. Proc. Natl Acad. Sci. USA 81, 1391–1393 (1984).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  65. Alberts, B. Molecular Biology of the Cell 7th edn (W. W. Norton & Company, 2022).

  66. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    Article  CAS  PubMed  Google Scholar 

  67. Leal, L. G. Advanced Transport Phenomena: Fluid Mechanics and Convective Transport Processes (Cambridge Univ. Press, 2007).

  68. Tinevez, J. Y. et al. TrackMate: an open and extensible platform for single-particle tracking. Methods 115, 80–90 (2017).

    Article  CAS  PubMed  Google Scholar 

  69. Cai, L. H., Panyukov, S. & Rubinstein, M. Mobility of nonsticky nanoparticles in polymer liquids. Macromolecules 44, 7853–7863 (2011).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  70. Squires, T. M. & Mason, T. G. Fluid mechanics of microrheology. Annu. Rev. Fluid Mech. 42, 413–438 (2010).

    Article  ADS  Google Scholar 

  71. Kalwarczyk, T. et al. Motion of nanoprobes in complex liquids within the framework of the length-scale dependent viscosity model. Adv. Colloid Interfac. 223, 55–63 (2015).

    Article  CAS  Google Scholar 

  72. Lekkerkerker, H. N. W. & Tuinier, R. Colloids and the Depletion Interaction (Springer, 2011).

  73. König, I., Soranno, A., Nettels, D. & Schuler, B. Impact of in-cell and in-vitro crowding on the conformations and dynamics of an intrinsically disordered protein. Angew. Chem. Int. Ed. 60, 10724–10729 (2021).

    Article  Google Scholar 

  74. Dertinger, T. et al. Two-focus fluorescence correlation spectroscopy: a new tool for accurate and absolute diffusion measurements. Chem. Phys. Chem. 8, 433–443 (2007).

    Article  CAS  PubMed  Google Scholar 

  75. Hofmann, H. et al. Polymer scaling laws of unfolded and intrinsically disordered proteins quantified with single-molecule spectroscopy. Proc. Natl Acad. Sci. USA 109, 16155–16160 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  76. Gilboa, B. et al. Confinement-free wide-field ratiometric tracking of single fluorescent molecules. Biophys. J. 117, 2141–2153 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  77. Armstrong, J. K., Wenby, R. B., Meiselman, H. J. & Fisher, T. C. The hydrodynamic radii of macromolecules and their effect on red blood cell aggregation. Biophys. J. 87, 4259–4270 (2004).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  78. Pamies, R., Cifre, J. G. H., Martinez, M. D. L. & de la Torre, J. G. Determination of intrinsic viscosities of macromolecules and nanoparticles. Comparison of single-point and dilution procedures. Colloid Polym. Sci. 286, 1223–1231 (2008).

    Article  CAS  Google Scholar 

  79. Gopich, I. V., Nettels, D., Schuler, B. & Szabo, A. Protein dynamics from single-molecule fluorescence intensity correlation functions. J. Chem. Phys. 131, 095102 (2009).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  80. Nettels, D., Gopich, I. V., Hoffmann, A. & Schuler, B. Ultrafast dynamics of protein collapse from single-molecule photon statistics. Proc. Natl Acad. Sci. USA 104, 2655–2660 (2007).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  81. Loman, A., Gregor, I., Stutz, C., Mund, M. & Enderlein, J. Measuring rotational diffusion of macromolecules by fluorescence correlation spectroscopy. Photochem. Photobiol. Sci. 9, 627–636 (2010).

    Article  CAS  PubMed  Google Scholar 

  82. Barth, A. et al. Unraveling multi-state molecular dynamics in single-molecule FRET experiments. I. Theory of FRET-lines. J. Chem. Phys. 156, 141501 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  83. Gopich, I. V. & Szabo, A. Theory of the energy transfer efficiency and fluorescence lifetime distribution in single-molecule FRET. Proc. Natl Acad. Sci. USA 109, 7747–7752 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  84. Koshioka, M., Sasaki, K. & Masuhara, H. Time-dependent fluorescence depolarization analysis in 3-dimensional microspectroscopy. Appl. Spect. 49, 224–228 (1995).

    Article  ADS  CAS  Google Scholar 

  85. Lindorff-Larsen, K. et al. Improved side-chain torsion potentials for the Amber ff99SB protein force field. Proteins 78, 1950–1958 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Luo, Y. & Roux, B. Simulation of osmotic pressure in concentrated aqueous salt solutions. J. Phys. Chem. Lett. 1, 183–189 (2010).

    Article  CAS  Google Scholar 

  87. Bussi, G., Donadio, D. & Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 126, 014101 (2007).

    Article  ADS  PubMed  Google Scholar 

  88. Parrinello, M. & Rahman, A. Polymorphic transitions in single crystals: a new molecular dynamics method. J. Appl. Phys. 52, 7182–7190 (1981).

    Article  ADS  CAS  Google Scholar 

  89. Darden, T., York, D. & Pedersen, L. Particle mesh Ewald: an Nlog(N) method for Ewald sums in large systems. J. Chem. Phys. 98, 10089–10092 (1993).

    Article  ADS  CAS  Google Scholar 

  90. Hess, B., Bekker, H., Berendsen, H. J. C. & Fraaije, J. G. E. M. LINCS: a linear constraint solver for molecular simulations. J. Comp. Chem. 18, 1463–1472 (1997).

    Article  CAS  Google Scholar 

  91. Abraham, M. J. et al. GROMACS: high performance molecular simulations through multi-levelvparallelism from laptops to supercomputers. SoftwareX 1-2, 19–25 (2015).

    Article  ADS  Google Scholar 

  92. Karanicolas, J. & Brooks, C. L. The origins of asymmetry in the folding transition states of protein L and protein G. Prot. Sci. 11, 2351–2361 (2002).

    Article  CAS  Google Scholar 

  93. Creighton, T. E. Proteins: Structures and Molecular Properties 2nd edn (W.H. Freeman, 1993).

  94. Tribello, G. A., Bonomi, M., Branduardi, D., Camilloni, C. & Bussi, G. PLUMED 2: new feathers for an old bird. Comput. Phys. Commun. 185, 604–613 (2014).

    Article  ADS  CAS  Google Scholar 

  95. Martinsen, J. H. et al. Structure, dynamics, and stability of the globular domain of human linker histone H1.0 and the role of positive charges. Protein Sci. 31, 918–932 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Rotkiewicz, P. & Skolnick, J. Fast procedure for reconstruction of full-atom protein models from reduced representations. J. Comput. Chem. 29, 1460–1465 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Vitalis, A. & Pappu, R. V. ABSINTH: a new continuum solvation model for simulations of polypeptides in aqueous solutions. J. Comput. Chem. 30, 673–699 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Holmstrom, E. D. et al. Accurate transfer efficiencies, distance distributions, and ensembles of unfolded and intrinsically disordered proteins from single-molecule FRET. Methods Enzymol. 611, 287–325 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Best, R. et al. Effect of flexibility and cis residues in single molecule FRET studies of polyproline. Proc. Natl Acad. Sci. USA 104, 18964–18969 (2007).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  100. Best, R. B., Hummer, G. & Eaton, W. A. Native contacts determine protein folding mechanisms in atomistic simulations. Proc. Natl Acad. Sci. USA 110, 17874–17879 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  101. Hasimoto, H. On the periodic fundamental solutions of the Stokes equations and their application to viscous flow past a cubic array of spheres. J. Fluid Mech. 5, 317–328 (1959).

    Article  ADS  MathSciNet  MATH  Google Scholar 

  102. Zhou, H.-X. Shape recovery of deformed biomolecular droplets: Dependence on condensate viscoelasticity. J. Chem. Phys. 155, 145102 (2021).

  103. Jeon, B. J. et al. Salt-dependent properties of a coacervate-like, self-assembled DNA liquid. Soft Matter 14, 7009–7015 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  104. Kask, B., Piksarv, P., Pooga, M., Mets, Ü. & Lippmaa E. Separation of the rotational contribution in fluorescence correlation experiments. Biophys. J. 55, 213–220 (1989).

  105. Lipari, G. & Szabo, A. Effect of librational motion on fluorescence depolarization and nuclear magnetic-resonance relaxation in macromolecules and membranes. Biophys. J. 30, 489–506 (1980).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  106. Boeynaems, S., De Decker, M., Tompa, P. & Van Den Bosch, L. Arginine-rich peptides can actively mediate liquid-liquid phase separation. Bio-Protocol 7, e2525 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  107. Qamar, S. et al. FUS phase separation is modulated by a molecular chaperone and methylation of arginine cation-pi interactions. Cell 173, 720–734 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Paloni, M., Bussi, G. & Barducci, A. Arginine multivalency stabilizes protein/RNA condensates. Protein Sci. 30, 1418–1426 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Vernon, R. M. et al. Pi-Pi contacts are an overlooked protein feature relevant to phase separation. eLife 7, e31486 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Peng, B. & Muthukumar, M. Modeling competitive substitution in a polyelectrolyte complex. J. Chem. Phys. 143, 243133 (2015).

  112. Sokolov, I. M. Models of anomalous diffusion in crowded environments. Soft Matter 8, 9043–9052 (2012).

    Article  ADS  CAS  Google Scholar 

  113. Doi, M. & Edwards, S. F. The Theory of Polymer Dynamics (Oxford Univ. Press, 1988).

Download references

Acknowledgements

We thank P. Banerjee, A. Borgia, M. Borgia, K. Ghosh, H. Hofmann, R. Pappu, A. Soranno, I. Sokolov, R. Tuinier and H.-X. Zhou for helpful discussions and comments on the manuscript; R. Zhu and P. Łukijańczuk for technical assistance in protein preparation; F. Sturzenegger for help with optical tweezer measurements and E. Lemke for providing the pBAD-Int-CBD-12His plasmid. This work was supported by the Swiss National Science Foundation (B.S.), the Novo Nordisk Foundation Challenge programme REPIN (grant no. NNF18OC0033926 to B.S.), the Intramural Research Program of the National Institute of Diabetes and Digestive and Kidney Diseases at the National Institutes of Health (R.B.B.), the Forschungskredit of the University of Zurich (N.G.) and the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement ID no. 898228 (A.C.). We used the computational resources of Piz Daint and Eiger at the CSCS Swiss National Supercomputing Center, and of the National Institutes of Health HPC Biowulf cluster (http://hpc.nih.gov). Mass spectrometry was performed at the Functional Genomics Center Zurich. FRAP and bead tracking were performed with support of the Center for Microscopy and Image Analysis, University of Zurich.

Author information

Authors and Affiliations

Authors

Contributions

N.G., M.T.I., R.B.B. and B.S. conceived the study. N.G. performed single-molecule experiments, FCS and microrheology. A.C. and N.G. characterized binodal curves. A.C. performed turbidity experiments. A.C., A.S. and M.F.N. performed protein purification and/or labelling. D.N. developed single-molecule instrumentation. N.G. and D.N. developed analysis tools for experimental data. N.G. analysed the experimental data, with input from A.C., A.S., B.S., D.N. and M.F.N. M.T.I. performed and analysed the simulations with the help of R.B.B. and input from B.S., N.G. and D.N. B.S. and R.B.B. supervised the project. N.G. and M.T.I. prepared the figures. B.S., N.G., M.T.I. and R.B.B. wrote the manuscript with contributions from all authors.

Corresponding authors

Correspondence to Nicola Galvanetto, Miloš T. Ivanović, Robert B. Best or Benjamin Schuler.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature thanks Marina Guenza, Ashok Deniz, Jörg Enderlein and Ziqing Winston Zhao for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

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

Extended data figures and tables

Extended Data Fig. 1 Dependence of phase separation on solution conditions and droplet fusion dynamics.

a. Phase separation is most pronounced in a charged-balanced mixture of H1 and ProTα. The extent of droplet formation was assessed using turbidity at 350 nm in TEK buffer with 50 mM KCl and at 120 mM KCl at a constant concentration of H1 (10 μM and 20 μM, at 50 mM and 120 mM KCl, respectively) and varying amounts of ProTα. At both salt concentrations, maximum phase separation was observed at a stoichiometric ratio of 1.2:1 for ProTα:H1, where the charges of the two proteins balance. b. Lohman-Record plot33 of the ionic strength dependence of the dilute (cdilute) over dense-phase protein concentration (cdense). If we treat the ratio cdilute/cdense as an effective equilibrium constant for the partitioning of H1 and ProTα between the dilute and dense phases, its logarithm approximates the free energy difference between the heterodimer in the dilute phase and in the dense phase. The slope of a graph of these values versus the logarithm of the ionic strength (or salt concentration) can then be interpreted in terms of the number of ions released33 upon the transfer of a ProTα–H1 dimer into the dense phase (since Log(cdilute/cdense) diverges close to the critical point, we only included data points up to 120 mM KCl). The resulting value of 2.5±0.7 released ions (uncertainty from error of the fit) is small compared to the ~18 ions released upon ProTα–H1 dimerization9,28, in accord with the small number of additional charge-charge interactions of ProTα in the dense phase compared to the heterodimer obtained from the simulations (Fig. 3e). Note that cdilute = 35 ± 5 μM at an ionic strength of 165 mM, which explains why no phase separation was observed in the NMR experiments of ProTα and H1 reported previously9. Even at the highest protein concentrations used there, the signal is expected to be dominated by the dilute phase, and in case droplets did form, their volume fraction was presumably too small to be apparent by eye. We chose to work at an ionic strength of 128 mM in the present work as a compromise between physiologically relevant salt concentrations and experimental feasibility, especially regarding sample consumption. c. Droplet relaxation upon droplet fusion (measured in dual-trap optical tweezers8, Fig. 1c) is single-exponential102, and the relaxation time is proportional to the radius of the final droplet, which indicates that the viscoelasticity of the dense phase on the millisecond timescale is dominated by the viscous (rather than the elastic) component8. In this case, the slope of the fit (dashed line) is67,103 \((2\lambda +3)(19\lambda +16)/[40(\lambda +1)]\cdot {\eta }_{{\rm{s}}}/\sigma \), where \(\lambda ={\eta }_{{\rm{m}}}/{\eta }_{{\rm{s}}}\) is the ratio of macroscopic (or bulk) viscosity in the droplet over the solvent viscosity (ηs = 0.001 Pa s), and σ is the interfacial tension. With the resulting value of 2.4·103 s/m for the slope and ηm = 0.3 Pa s, we estimate σ ≈ 1.2·10−4 N/m.

Source Data

Extended Data Fig. 2 The ProTα–H1 dimer is the dominant population in the dilute phase.

Single-molecule transfer efficiency histogram of ProTαC (labeled at position 56 and 110) in the dilute phase at 128 mM ionic strength (TEK buffer including 120 mM KCl). The phase-separated mixture was centrifuged, so that the dense phase coalesced into a single large droplet and no small droplets remained in the dilute phase. The dilute phase was aspirated and transferred into a sample chamber for single-molecule measurements. In the fit (lines), the centers of the Gaussian peak functions were constrained to the transfer efficiencies measured for unbound ProTα and the ProTα–H1 dimer (Fig. 2f) to within experimental uncertainty. The shaded peak near a transfer efficiency of zero originates from molecules lacking an active acceptor dye.

Source Data

Extended Data Fig. 3 Polarization-resolved fluorescence probing rotational effects.

a. Donor and acceptor emission autocorrelations (green and red, respectively; parallel vs perpendicular channels) and donor-acceptor crosscorrelation (purple; sum of correlations of respective parallel and perpendicular channels) of the FRET-active subpopulation of labeled ProTαC in the dense phase when a polarizing beam splitter is used show an asymmetry of the branches for positive and negative lag-times, τ, in the positively correlated component (correlation time of 30 ns). In contrast, this component is more symmetric when a 50-50 beam splitter is used (b), indicating that the component is caused by residual polarization anisotropy104. (c–h) Time-resolved anisotropy decays, r(t), measured for double-labeled ProTαC unbound (c,d), in the dimer (e,f), and in the dense phase (g,h) with pulsed interleaved excitation using (c,e,g) photons from donor-only bursts (transfer efficiency < 0.1, excitation at 532 nm) or (d,f,h) acceptor photons from bursts with transfer efficiency >0.2 (excitation at 635 nm). Data were fitted with the function \(r(t)={r}_{0}((1-{A}_{{\rm{s}}{\rm{l}}{\rm{o}}{\rm{w}}}){e}^{-t/{\tau }_{{\rm{f}}{\rm{a}}{\rm{s}}{\rm{t}}}}+{A}_{{\rm{s}}{\rm{l}}{\rm{o}}{\rm{w}}}){e}^{-t/{\tau }_{{\rm{s}}{\rm{l}}{\rm{o}}{\rm{w}}}}\) (dashed black lines)105 with \({r}_{0}\) = 0.4. No significant amplitude \({A}_{{\rm{s}}{\rm{l}}{\rm{o}}{\rm{w}}}\) for a slow component is present for free ProTαC (c, d), and only a small amplitude in the dimer (e,f). In the dense phase (g,h), a distinct slow decay component is observed in the anisotropy decay, which is well described with the decay time τslow = 30 ns from the correlated component of the nsFCS (a,b). This agreement further supports the role of residual rotation as the source of the latter. (i,j) Time-resolved anisotropy decays for free Cy3B in the dilute (i) and dense phase (j). The dilute-phase decay was fit with a single exponential, \(r\left(t\right)={r}_{0}\,{e}^{-t/\tau }\), and the resulting value of \(\tau =0.53\) ns was used to obtain the hydrodynamic radius of Cy3B based on the Stokes-Einstein-Debye relation, \(\tau =({\eta }_{{\rm{e}}{\rm{f}}{\rm{f}}}\,\frac{4}{3}\pi {R}_{{\rm{p}}{\rm{r}}{\rm{o}}{\rm{b}}{\rm{e}}}^{3})/({k}_{{\rm{B}}}T)\). With the viscosity of water (\(0.0010\,{\rm{P}}{\rm{a}}\,{\rm{s}}\)), we obtained 0.80 nm for the radius of Cy3B, within the range of the previously reported values (0.76 ± 0.04 nm)76. (j) The anisotropy decay in the dense phase was fit with a sum of two exponentials, \(r(t)={r}_{0}((1-{A}_{{\rm{s}}{\rm{l}}{\rm{o}}{\rm{w}}}){e}^{-t/{\tau }_{{\rm{f}}{\rm{a}}{\rm{s}}{\rm{t}}}}+{A}_{{\rm{s}}{\rm{l}}{\rm{o}}{\rm{w}}}\,{e}^{-t/{\tau }_{{\rm{s}}{\rm{l}}{\rm{o}}{\rm{w}}}})\). The effective viscosities obtained by means of the Stokes-Einstein-Debye relation from the fast and slow components, \({\tau }_{{fast}}\) and \({\tau }_{{slow}}\), are reported in Fig. 1e, and we assign the fast component to the rotational diffusion of the dye virtually unaffected by attractive protein interactions. Note that despite the slow rotational component of Cy3B, almost no partitioning of the dye into the droplets was observed (partition constant <1.05 from confocal fluorescence microscopy images).

Source Data

Extended Data Fig. 4 ProTα labeled at positions 2 and 56 (ProTαN) shows behaviour similar to ProTα labeled at positions 56 and 110 (ProTαC, Fig. 2).

a. Single-molecule transfer efficiency histograms of ProTαN at 128 mM ionic strength (TEK buffer including 120 mM KCl) as a monomer free in solution (top), in the 1:1 complex with H1 (middle), and within droplets (bottom) measured with continuous-wave excitation. Note the greater compaction in the dense phase compared to the ProTα–H1 dimer than for ProTαC. b. 2D histograms of relative donor (above diagonal) and acceptor fluorescence lifetimes (below diagonal) versus FRET efficiency for all detected bursts measured with pulsed excitation of ProTαN. The straight line shows the dependence expected for fluorophores separated by a static distance; curved lines show the dependences for fluorophores that rapidly sample a distribution of distances (self-avoiding walk (SAW-ν)40, see Methods; upper line: donor lifetime; lower line: acceptor lifetime). c. nsFCS probing chain dynamics based on intramolecular FRET in double-labeled ProTαN; data show donor–acceptor fluorescence cross-correlations with fits (black lines). Reconfiguration times, τr, are averages of n = 3 independent measurements (uncertainties discussed in Methods).

Source Data

Extended Data Fig. 5 Full FCS curves with logarithmic binning.

Donor and acceptor autocorrelations (green, red) and donor-acceptor crosscorrelations (purple; same color scheme as in Fig. 2h, which shows the same data and fits but on a linear scale and normalized to an amplitude of 1 at ±3 μs) of ProTαC (labeled at position 56 and 110) at 128 mM ionic strength (TEK buffer with 120 mM KCl) as an unbound monomer in solution (a), in the 1:1 complex with H1 (b), and within ProTα–H1 droplets (c). For each sample, the three correlations are fitted globally (black solid lines, see Methods) with shared correlation times for translational diffusion (τD), triplet blinking (τT), dye rotation (τrot), and conformational dynamics (τcd); photon antibunching (τab) is fitted individually. τcd was then converted to the reconfiguration time of the chain, τr, as previously described79 (we note that the conversion from τcd to τr does not entail a large change in timescale, and τcd and τr differ by less than 20% in all cases investigated here). τD, τT, τrot, τr, and τab are shown in the panels if the corresponding term was included in the fit function (Eq. 6), and they point to their respective timescales. The value of τr reported here is the mean of three measurements, as in Fig. 2h, and corresponds to the distance correlation time between the dyes at position 56 and 110.79 τT in the donor-acceptor cross correlation in (B) shows a small negative amplitude, possibly indicating a slight contribution of slower distance dynamics on the microsecond timescale. Note that the deviation between fit and measurement in (c) for the translational diffusion component is caused by sample scanning, which was required to improve statistics inside the droplets.

Source Data

Extended Data Fig. 6 Composition of dense versus dilute phase from simulations.

Mass concentrations of protein, water, all components (protein, water, and ions; left), and number density of ions (right) along the z axis of the simulation box (see inset on the right). The water mass concentration in the dense phase (central part of the slab, 15 nm < z < 30 nm) is ~80.7% of the water concentration in the bulk regions (z < 2.5 nm and z > 40.0 nm). The number density of ions in the dense phase (15 nm < z < 30 nm) is ~88.4% of the value close to the box edges (z < 1.5 nm and z > 41.0 nm). With respect to only the water content in the respective phases, the ion concentration is ~10% higher in the dense phase than in the dilute phase.

Source Data

Extended Data Fig. 7 Equilibration of dense phase simulation, stability of H1 globular domain in simulations, and robustness to cutoff variation.

a. Protein density in the central part of the slab simulation as a function of time, calculated in 50-ns blocks. The first 1.5 µs of the simulation (shaded band) were treated as equilibration and omitted from further analysis. b. Stability of the H1 globular domains (GDs), quantified as the backbone RMSD between simulated and experimental structure (PDB 6HQ1)95, over the course of dimer (left) and dense-phase simulations (middle). The fraction of partially unfolded domains (< 10% with RMSD > 0.4 nm) is in line with the experimental stability previously determined in dilute solution95. Note that the backbone RMSD of 0.2-0.4 nm for the folded domain can be attributed to the flexibility of the loops in the structure, illustrated by superposition of two structures with RMSD = 0.4 nm (right). c. Histograms of the number of H1 chains simultaneously interacting with a single ProTα chain (left) and vice versa (middle) using different distance cutoffs (see legend; 1.0 nm91 was used in Fig. 3c). Note that the number of ProTα chains interacting with a single H1 chain is always ~1.2 times the number of H1 chains interacting with a single ProTα chain (right), as expected from charge balance (Extended Data Fig. 1a). d. Average number of contacts that each residue of ProTα makes in the heterodimer with H1 (grey) and in the dense phase (purple) with different distance cutoffs (1.0 nm — approximately the Bjerrum length — was used in the Fig. 3e). As expected, the increase in the number of contacts with the cutoff is more pronounced in the dense phase than in the dimer, reflecting the higher protein density in the condensate. Owing to the computational costs of the distance calculations for each residue (10,752 distances in total), only 500 ns of the dense-phase trajectory (2.0 to 2.5 μs) were used for cutoff variation (in contrast to Fig. 3e, where the complete trajectory was used). e. Distribution of the lifetimes of contacts formed by ProTα residues in the dimer (grey) and in the dense phase (purple) using different upper bounds for the contact definition (see Methods; 0.8 nm100 was used in Fig. 3f). Owing to the computational costs of the lifetime calculations, only 500 ns of the dense-phase trajectory (1.5 to 2.0 μs) and only one dimer simulation was used for cutoff variation (in contrast to Fig. 3f, where the complete trajectory and all dimer simulations were used). While the increased cutoff leads to a slight increase in the absolute values of the lifetimes, the reported trends are consistent: the distributions of longer-lived contacts are very similar for the dimer and dense-phase simulations, and the number of short-lived contacts is larger in the dense phase. f. Root-mean-square displacement (RMSD) of the 112 individual ProTα residues with different contact lifetimes (see legend) vs their average frequency of contact formation.

Source Data

Extended Data Fig. 8 Contact lifetime heatmaps.

Average lifetime of residue-residue contacts from 6 simulations of the ProTα–H1 dimer (a) and the dense-phase simulation (b). Numbers along the bottom and left denote the residue numbers of ProTα and H1, respectively. Orange rectangles denote the globular domain (GD) of H1 (residues 22 to 96). Frequency of contacts (i.e. the number of newly formed contacts by one ProTα molecule per nanosecond) calculated from dimer and dense phase simulations are shown in (c), and (d), respectively. Blue and red bars at the top and on the right side of the plots denote positively and negatively charged residues of ProTα and H1, respectively. In general, the N-terminal part of ProTα makes fewer contacts than the rest of the chain both in the dimer and dense phase simulations (see also Fig 3e), and the lifetime of those contacts is on average shorter, especially in the dense-phase simulation. As is obvious from (d), contacts between oppositely charged residues are most frequent. White regions in a and c correspond to reside-residue combinations that were never formed during the simulations. White regions are particularly frequent in the GD, since it remains folded during the dimer simulations (Extended Data Fig. 7b). Some of the GD residues make relatively long-lived contacts, but those contacts are infrequent. In contrast to the dimer simulations, some residues of the GD do form contacts with ProTα residues in the dense phase simulation, since a small fraction of partially unfolded GDs are populated (Extended Data Fig. 7b), as expected from the low equilibrium stability of the GD9,95. e–h. Residue type-specific contact lifetime heatmaps. Average lifetimes of residue-residue contacts in the ProTα–H1 dimer (e) and the dense-phase simulations (f) classified by residue types. Excess population of contacts for specific residue pairs in the ProTα–H1 dimer (g) and in the dense-phase simulation (h) (see Methods for details). Residue pairs that are never observed (white squares) and extremely long-lived pairs (dark blue) in (e) correspond to residue types that are infrequent in the ProTα and H1 sequence (compare with i). In the dense phase, Arg forms contacts that are on average longer-lived than any other residue (f), in line with the phase separation-promoting role of Arg106,107,108,109,110. The excess populations (see Methods) of contacts for specific residue pairs suggest that the interactions between charged residues are the most favorable interactions both in the dimer and in the dense-phase simulations. Note that the oppositely charged residues Glu (most abundant residue in ProTα) and Lys (most abundant residue in H1) form the largest number of contacts (g,h) but have lifetimes comparable to other residue pairs (e,f). i. Number of each type of residues in ProTα (red) and H1 (blue).

Source Data

Extended Data Fig. 9 Lifetimes of non-attractive collisional contacts and competitive substitution between residues.

a. The duration of a contact between two residues was estimated from the time when the distance between any two heavy atoms of the two residues falls below 0.38 nm to the first time when no distance between any two heavy atoms of those residues is below 0.80 nm (see Methods). We used the transition path times of residue–residue contact breaking as an estimate for the lifetime of non-attractive collisional contacts between two residues. The transition path time for the breaking of a given contact was estimated as the time from the last time when the distance between any two heavy atoms of the two residues is below 0.38 nm to the first time it reaches 0.8 nm. The timescale expected for non-attractive collisions in the dense-phase simulation (shaded area in Fig. 3f) was estimated as the time that includes 95% of all transition path times in the dense-phase simulation. b. Comparison between the contact lifetimes and the transition path times in the ProTα–H1 dimer and the dense phase (see legend). The areas under the curves correspond to the total numbers of contact events per chain per nanosecond. c. A fingerprint of rapid exchange or competitive substitution111 between charged side chains in the dense phase. Average number of contacts at the time when the contact between two residues is broken plotted as a function of the average number of contacts that those two residues make with other residues during the time being in contact. Given the large number of contact events in the dense phase simulation, only every 20,000th data point is plotted. The definition of a contact is identical to the one described in Methods, but the average number of contacts per residue is larger than the one shown in Fig. 3e since in this case the bonds between neighboring residues were also recorded as contacts. The significantly lower value of the slope of a linear fit in the dimer simulation suggests that multiple contacts tend to be broken simultaneously in this case owing to the concerted motions of parts of the protein chains. In contrast, owing to the high local density of potential interaction partners in the dense phase and the competition for contacts, less contacts are broken simultaneously, as the interaction partners are often rapidly substituted (Fig. 3h), resulting in the greater slope in the dense phase simulation.

Source Data

Extended Data Fig. 10 Mean square displacement (MSD) curves from molecular dynamics simualtions simulations reveal subdiffusion.

a. Center-of-mass diffusion of ProTα in the dense phase (purple, average of 96 chains) compared to ProTα in the ProTα–H1 dimer (grey, average of 6 chains). In the dimer, at all timescales investigated, the diffusion of ProTα is Brownian, whereas in the dense phase, we observe subdiffusive behavior at timescales equal to or shorter than the chain reconfiguration time (shaded bands indicate full-length chain reconfiguration time ± uncertainty), as expected in the presence of cooperative dynamics of the network48 (MSDs are only shown for the time range where the standard deviation σMSD < 0.5·MSD). b,c. Comparison between the diffusion of residue 1 of ProTα, of the central residue 58, and of the ProTα center of mass in the dimer (b) and the dense phase (c). The residues of an ideal chain are expected to show subdiffusive behavior in a time window between tKuhn, the time a residue needs to diffuse over the Kuhn length of the chain, and the time the entire chain takes to diffuse a distance corresponding to its own size112, which, for a Rouse chain113, approximately corresponds to the chain reconfiguration time, τr. Below tKuhn, the individual residues are expected to diffuse independently of the chain. Building on the ideal chain model, in (f) we report the diffusion exponent for times below 2 ns (approximately tKuhn), where the single-residue behavior is largely unaffected by the slowdown due to chain reconfiguration. d,e. Same data as in (b,c), but in linear scale to highlight the transition at timescales >τr, where the diffusion of the entire chain dominates the diffusion of the individual residues. The yellow and orange vertical lines indicate the MSD travelled by the residue in excess of the MSD of the center of mass of the chain. Dashed lines indicate the slope expected for Brownian dynamics. f. Diffusion of individual ProTα residues (1–112) is examined in terms of their mean squared displacement, MSD(t) = 6Dtα, for timescales shorter than tKuhn (see b,c), where D is the diffusion coefficient, t is the lag time, and α = 1 for Brownian diffusion. Diffusion of the residues in the ProTα–H1 dimer is close to Brownian and does not correlate with the average contact lifetime of the corresponding residues, whereas in the dense phase, the diffusion of the residues is more subdiffusive (α < 1) and shows a negative correlation with their average contact lifetime. The residues in the dense phase with low average contact lifetime show less subdiffusive behavior but form a larger number of contacts per unit time (compare with Fig. 3g).

Source Data

Extended Data Table 1 Amino acid sequences of the proteins used

Supplementary information

Reporting Summary

Peer Review File

Supplementary Video 1

All-atom explicit-solvent simulation of ProTα–H1 dense phase. One ProTα chain is highlighted in red (chain 59) and four interacting H1 chains in different shades of blue. Other surrounding ProTα and H1 chains are shown semitransparently in red and blue, respectively. The image is centred on the COM of the highlighted ProTα chain. The total length of the simulation is 6 μs, and the first 1.5 μs were omitted from the analysis and the video; the video is shown at 2 ns per frame. To slightly smooth the motion, a filter with a time constant of 4 ns was applied to all frames. Protein hydrogen atoms, water molecules and ions are omitted for clarity.

Supplementary Video 2

All-atom explicit-solvent simulation of ProTα–H1 dense phase. One ProTα chain is highlighted in red (chain 60) and four interacting H1 chains in different shades of blue. Other surrounding ProTα and H1 chains are shown semitransparently in red and blue, respectively. The image is centred on the COM of the highlighted ProTα chain. The total length of the simulation is 6 μs, and the first 1.5 μs were omitted from the analysis and the video; the video is shown at 2 ns per frame. To slightly smooth the motion, a filter with a time constant of 4 ns was applied to all frames. Protein hydrogen atoms, water molecules and ions are omitted for clarity.

Supplementary Video 3

Example of protein dynamics at short timescales in the dense phase. One ProTα chain is highlighted in red and four interacting H1 chains in different shades of blue. Other surrounding ProTα and H1 chains are shown semitransparently in red and blue, respectively. The image is centred on the COM of the highlighted ProTα chain; the video is shown at 100 ps per frame. To slightly smooth the motion, a filter with a time constant of 200 ps was applied to all frames. Protein hydrogen atoms, water molecules and ions are omitted for clarity.

Source data

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

Galvanetto, N., Ivanović, M.T., Chowdhury, A. et al. Extreme dynamics in a biomolecular condensate. Nature 619, 876–883 (2023). https://doi.org/10.1038/s41586-023-06329-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-023-06329-5

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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