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:

Molecular interactions contributing to FUS SYGQ LC-RGG phase separation and co-partitioning with RNA polymerase II heptads

Abstract

The RNA-binding protein FUS (Fused in Sarcoma) mediates phase separation in biomolecular condensates and functions in transcription by clustering with RNA polymerase II. Specific contact residues and interaction modes formed by FUS and the C-terminal heptad repeats of RNA polymerase II (CTD) have been suggested but not probed directly. Here we show how RGG domains contribute to phase separation with the FUS N-terminal low-complexity domain (SYGQ LC) and RNA polymerase II CTD. Using NMR spectroscopy and molecular simulations, we demonstrate that many residue types, not solely arginine-tyrosine pairs, form condensed-phase contacts via several interaction modes including, but not only sp2-π and cation-π interactions. In phases also containing RNA polymerase II CTD, many residue types form contacts, including both cation-π and hydrogen-bonding interactions formed by the conserved human CTD lysines. Hence, our data suggest a surprisingly broad array of residue types and modes explain co-phase separation of FUS and RNA polymerase II.

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: Domains outside the SYGQ-rich low-complexity domain contribute to FUS LLPS.
Fig. 2: The RGG3 and SYGQ LC remain disordered in a two-component condensed phase.
Fig. 3: Interactions between SYGQ LC and RGG3 in the condensed phase involve multiple residue pairs.
Fig. 4: Diverse interaction modes contribute to FUS SYGQ LC and RGG3 interactions within the condensed phase.
Fig. 5: RGG domains interact with RNA polymerase II CTD tail.
Fig. 6: Lysine residues within RNA polymerase II CTD contribute to interactions with FUS within the condensed phase.
Fig. 7: Interaction modes and impact of CTD lysine residues on RNA polymerase II CTD and FUS domain interactions.
Fig. 8: Multiple residue types and interaction modes contribute to multicomponent condensed phases of FET proteins and RNA polymerase II.

Similar content being viewed by others

Data availability

Chemical shift assignments for the RGG domains can be accessed using the BMRB accession 51067, 51068, 51069. Raw NMR data files can be found at https://doi.org/10.6084/m9.figshare.16598861. Source data are provided with this paper. All other data are available from the corresponding author upon reasonable request.

Code availability

Simulation software described in Methods section are publicly available and can be found at http://www.gromacs.org/ for the atomistic resolution simulations.

References

  1. Brangwynne, C. P., Mitchison, T. J. & Hyman, A. A. Active liquid-like behavior of nucleoli determines their size and shape in Xenopus laevis oocytes. Proc. Natl Acad. Sci. USA 108, 4334–4339 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Feric, M. et al. Coexisting liquid phases underlie nucleolar subcompartments. Cell 165, 1686–1697 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  5. Boija, A. et al. Transcription factors activate genes through the phase-separation capacity of their activation domains. Cell 175, 1842–1855.e16 (2018).

    Article  CAS  PubMed  Google Scholar 

  6. Cho, W. K. et al. Mediator and RNA polymerase II clusters associate in transcription-dependent condensates. Science 361, 412–415 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  8. Svetoni, F., Frisone, P. & Paronetto, M. P. Role of FET proteins in neurodegenerative disorders. RNA Biol. 13, 1089–1102 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Hoell, J. I. et al. RNA targets of wild-type and mutant FET family proteins. Nat. Struct. Mol. Biol. 18, 1428–1431 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Kapeli, K. et al. Distinct and shared functions of ALS-associated proteins TDP-43, FUS and TAF15 revealed by multisystem analyses. Nat. Commun. 7, 12143 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Loughlin, F. E. et al. The solution structure of FUS bound to RNA reveals a bipartite mode of RNA recognition with both sequence and shape specificity. Mol. Cell 73, 490–504.e6 (2019).

    Article  CAS  PubMed  Google Scholar 

  12. Schwartz, J. C., Wang, X., Podell, E. R. & Cech, T. R. RNA seeds higher-order assembly of FUS protein. Cell Rep. 5, 918–925 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Shelkovnikova, T. A., Robinson, H. K., Southcombe, J. A., Ninkina, N. & Buchman, V. L. Multistep process of FUS aggregation in the cell cytoplasm involves RNA-dependent and RNA-independent mechanisms. Hum. Mol. Genet. 23, 5211–5226 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Burke, K. A., Janke, A. M., Rhine, C. L. & Fawzi, N. L. Residue-by-residue view of in vitro FUS granules that bind the C-terminal domain of RNA polymerase II. Mol. Cell 60, 231–241 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Kato, M. et al. Cell-free formation of RNA granules: low complexity sequence domains form dynamic fibers within hydrogels. Cell 149, 753–767 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Murakami, T. et al. ALS/FTD mutation-induced phase transition of FUS liquid droplets and reversible hydrogels into irreversible hydrogels impairs RNP granule function. Neuron 88, 678–690 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Sun, Z. et al. Molecular determinants and genetic modifiers of aggregation and toxicity for the ALS disease protein FUS/TLS. PLoS Biol. 9, e1000614 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Patel, A. et al. A liquid-to-solid phase transition of the ALS protein FUS accelerated by disease mutation. Cell 162, 1066–1077 (2015).

    Article  CAS  PubMed  Google Scholar 

  19. Murthy, A. C. et al. Molecular interactions underlying liquid−liquid phase separation of the FUS low-complexity domain. Nat. Struct. Mol. Biol. 26, 637–648 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. 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  CAS  PubMed  PubMed Central  Google Scholar 

  21. Yoshizawa, T. et al. Nuclear import receptor inhibits phase separation of FUS through binding to multiple sites. Cell 173, 693–705 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Kang, J., Lim, L., Lu, Y. & Song, J. A unified mechanism for LLPS of ALS/FTLD-causing FUS as well as its modulation by ATP and oligonucleic acids. PLoS Biol. 17, e3000327 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  23. Qamar, S. et al. FUS phase separation is modulated by a molecular chaperone and methylation of arginine cation–π interactions. Cell 173, 720–734.e15 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Bogaert, E. et al. Molecular dissection of FUS points at synergistic effect of low-complexity domains in toxicity. Cell Rep. 24, 529–537.e4 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Hofweber, M. et al. Phase separation of FUS is suppressed by its nuclear import receptor and arginine methylation. Cell 173, 706–719.e13 (2018).

    Article  CAS  PubMed  Google Scholar 

  26. Wang, J. et al. A molecular grammar governing the driving forces for phase separation of prion-like RNA binding proteins. Cell 174, 688–699.e16 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. 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 

  28. Schuster, B. S. et al. Identifying sequence perturbations to an intrinsically disordered protein that determine its phase-separation behavior. Proc. Natl Acad. Sci. USA 117, 11421–11431 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Bremer, A. et al. Deciphering how naturally occurring sequence features impact the phase behaviors of disordered prion-like domains. Preprint at bioRxiv https://doi.org/10.1101/2021.01.01.425046 (2021).

  30. Schwartz, J. C. et al. FUS binds the CTD of RNA polymerase II and regulates its phosphorylation at Ser2. Genes Dev. 26, 2690–2695 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Kwon, I. et al. Phosphorylation-regulated binding of RNA polymerase II to fibrous polymers of low-complexity domains. Cell 155, 1049 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Zinszner, H., Albalat, R. & Ron, D. A novel effector domain from the RNA-binding protein TLS or EWS is required for oncogenic transformation by CHOP. Genes Dev. 8, 2513–2526 (1994).

    Article  CAS  PubMed  Google Scholar 

  33. Ichikawa, H., Shimizu, K., Hayashi, Y. & Ohki, M. An RNA-binding protein gene, TLS/FUS, is fused to ERG in human myeloid leukemia with t(16;21) chromosomal translocation. Cancer Res. 54, 2865–2868 (1994).

    CAS  PubMed  Google Scholar 

  34. Wei, M.-T. et al. Nucleated transcriptional condensates amplify gene expression. Nat. Cell Biol. 22, 1187–1196 (2020).

    Article  CAS  PubMed  Google Scholar 

  35. Lu, F., Portz, B. & Gilmour, D. S. The C-terminal domain of RNA polymerase II is a multivalent targeting sequence that supports Drosophila development with only consensus heptads. Mol. Cell 73, 1232–1242 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Boehning, M. et al. RNA polymerase II clustering through carboxy-terminal domain phase separation. Nat. Struct. Mol. Biol. 25, 833–840 (2018).

    Article  CAS  PubMed  Google Scholar 

  37. Monahan, Z. et al. Phosphorylation of the FUS low-complexity domain disrupts phase separation, aggregation, and toxicity. EMBO J. 36, 2951–2967 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Alshareedah, I. et al. Interplay between short-range attraction and long-range repulsion controls reentrant liquid condensation of ribonucleoprotein–RNA complexes. J. Am. Chem. Soc. 141, 14593–14602 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Banerjee, P. R., Milin, A. N., Moosa, M. M., Onuchic, P. L. & Deniz, A. A. Reentrant phase transition drives dynamic substructure formation in ribonucleoprotein droplets. Angew. Chem. Int. Ed. Engl. 56, 11354–11359 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Ukmar-Godec, T. et al. Lysine/RNA-interactions drive and regulate biomolecular condensation. Nat. Commun. 10, 2909 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  41. 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 

  42. Kay, L. E., Torchia, D. A. & Bax, A. Backbone dynamics of proteins as studied by 15N inverse detected heteronuclear NMR spectroscopy: application to staphylococcal nuclease. Biochemistry 28, 8972–8979 (1989).

    Article  CAS  PubMed  Google Scholar 

  43. Fawzi, N. L., Ying, J., Torchia, D. A. & Clore, G. M. Kinetics of amyloid β monomer-to-oligomer exchange by NMR relaxation. J. Am. Chem. Soc. 132, 9948–9951 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Neuhaus, D. & Williamson, M. P. The Nuclear Overhauser Effect in Structural and Conformational Analysis (Wiley, 2000).

    Google Scholar 

  45. 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 

  46. Chong, P. A., Vernon, R. M. & Forman-Kay, J. D. RGG/RG motif regions in RNA binding and phase separation. J. Mol. Biol. 430, 4650–4665 (2018).

    Article  CAS  PubMed  Google Scholar 

  47. Zerze, G. H., Best, R. B. & Mittal, J. Sequence- and temperature-dependent properties of unfolded and disordered proteins from atomistic simulations. J. Phys. Chem. B 119, 14622–14630 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Kjaergaard, M. et al. Temperature-dependent structural changes in intrinsically disordered proteins: formation of α-helices or loss of polyproline II? Protein Sci. 19, 1555–1564 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Wuttke, R. et al. Temperature-dependent solvation modulates the dimensions of disordered proteins. Proc. Natl Acad. Sci. USA 111, 5213–5218 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Janke, A. M. et al. Lysines in the RNA polymerase II C-terminal domain contribute to TAF15 fibril recruitment. Biochemistry 57, 2549–2563 (2018).

    Article  CAS  PubMed  Google Scholar 

  51. Fawzi, N. L. et al. Structure and dynamics of the Aβ21–30 peptide from the interplay of NMR experiments and molecular simulations. J. Am. Chem. Soc. 130, 6145–6158 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Nott, T. J. et al. Phase transition of a disordered nuage protein generates environmentally responsive membraneless organelles. Mol. Cell 57, 936–947 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Tsang, B. et al. Phosphoregulated FMRP phase separation models activity-dependent translation through bidirectional control of mRNA granule formation. Proc. Natl Acad. Sci. USA 116, 4218–4227 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Ryan, V. H. et al. Mechanistic view of hnRNPA2 low-complexity domain structure, interactions, and phase separation altered by mutation and arginine methylation. Mol. Cell 69, 465–479.e7 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Gibbs, E. B., Cook, E. C. & Showalter, S. A. Application of NMR to studies of intrinsically disordered proteins. Arch. Biochem. Biophys. 628, 57–70 (2017).

    Article  CAS  PubMed  Google Scholar 

  56. Gibbs, E., Perrone, B., Hassan, A., Kümmerle, R. & Kriwacki, R. NPM1 exhibits structural and dynamic heterogeneity upon phase separation with the p14ARF tumor suppressor. J. Magn. Reson. 310, 106646 (2020).

    Article  CAS  PubMed  Google Scholar 

  57. Reichheld, S. E., Muiznieks, L. D., Keeley, F. W. & Sharpe, S. Direct observation of structure and dynamics during phase separation of an elastomeric protein. Proc. Natl Acad. Sci. USA 114, E4408–E4415 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Kim, T. H. et al. Phospho-dependent phase separation of FMRP and CAPRIN1 recapitulates regulation of translation and deadenylation. Science 365, 825–829 (2019).

    Article  CAS  PubMed  Google Scholar 

  59. Kim, T. H. et al. Interaction hot spots for phase separation revealed by NMR studies of a CAPRIN1 condensed phase. Proc. Natl Acad. Sci. USA 118, e2014897118 (2021).

    Google Scholar 

  60. Wong, L. E., Kim, T. H., Muhandiram, D. R., Forman-Kay, J. D. & Kay, L. E. NMR experiments for studies of dilute and condensed protein phases: application to the phase-separating protein CAPRIN1. J. Am. Chem. Soc. 142, 2471–2489 (2020).

    Article  PubMed  Google Scholar 

  61. Dignon, G. L., Zheng, W., Kim, Y. C., Best, R. B. & Mittal, J. Sequence determinants of protein phase behavior from a coarse-grained model. PLoS Comput. Biol. 14, e1005941 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  62. Das, S., Lin, Y. H., Vernon, R. M., Forman-Kay, J. D. & Chan, H. S. Comparative roles of charge, π, and hydrophobic interactions in sequence-dependent phase separation of intrinsically disordered proteins. Proc. Natl Acad. Sci. USA 117, 28795–28805 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Paloni, M., Bailly, R., Ciandrini, L. & Barducci, A. Unraveling molecular interactions in liquid–liquid phase separation of disordered proteins by atomistic simulations. J. Phys. Chem. B 124, 9009–9016 (2020).

    Article  CAS  PubMed  Google Scholar 

  64. 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 

  65. Tang, W. S., Fawzi, N. L. & Mittal, J. Refining all-atom protein force fields for polar-rich, prion-like, low-complexity intrinsically disordered proteins. J. Phys. Chem. B 124, 9505–9512 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Zerze, G. H., Zheng, W., Best, R. B. & Mittal, J. Evolution of all-atom protein force fields to improve local and global properties. J. Phys. Chem. Lett. 10, 2227–2234 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. 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 

  68. Turupcu, A., Tirado-Rives, J. & Jorgensen, W. L. Explicit representation of cation−π interactions in force fields with 1/r4 nonbonded terms. J. Chem. Theory Comput. 16, 7184–7194 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  70. Ambadipudi, S., Reddy, J. G., Biernat, J., Mandelkow, E. & Zweckstetter, M. Residue-specific identification of liquid phase separation hot spots of the Alzheimer’s disease-related protein Tau. Chem. Sci. 10, 6503–6507 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Guo, Y. E. et al. Pol II phosphorylation regulates a switch between transcriptional and splicing condensates. Nature 572, 543–548 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Lu, H. et al. Phase-separation mechanism for C-terminal hyperphosphorylation of RNA polymerase II. Nature 558, 318–323 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Delaglio, F. et al. NMRPipe: a multidimensional spectral processing system based on UNIX pipes. J. Biomol. NMR 6, 277–293 (1995).

    Article  CAS  PubMed  Google Scholar 

  74. Lee, W., Tonelli, M. & Markley, J. L. NMRFAM-SPARKY: enhanced software for biomolecular NMR spectroscopy. Bioinformatics 31, 1325–1327 (2015).

    Article  PubMed  Google Scholar 

  75. Thiele, C. M., Petzold, K. & Schleucher, J. EASY ROESY: reliable cross-peak integration in adiabatic symmetrized ROESY. Chemistry 15, 585–588 (2009).

    Article  CAS  PubMed  Google Scholar 

  76. Hess, B., Kutzner, C., Van Der Spoel, D. & Lindahl, E. GRGMACS 4: algorithms for highly efficient, load-balanced, and scalable molecular simulation. J. Chem. Theory Comput. 4, 435–447 (2008).

    Article  CAS  PubMed  Google Scholar 

  77. 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  CAS  PubMed  Google Scholar 

  78. 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 

  79. Sugita, Y. & Okamoto, Y. Replica-exchange molecular dynamics method for protein folding. Chem. Phys. Lett. 314, 141–151 (1999).

    Article  CAS  Google Scholar 

  80. Bonomi, M. & Parrinello, M. Enhanced sampling in the well-tempered ensemble. Phys. Rev. Lett. 104, 190601 (2010).

    Article  CAS  PubMed  Google Scholar 

  81. Barducci, A., Bussi, G. & Parrinello, M. Well-tempered metadynamics: a smoothly converging and tunable free-energy method. Phys. Rev. Lett. 100, 020603 (2008).

    Article  PubMed  Google Scholar 

  82. 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 

  83. Brooks, B. R. et al. CHARMM: the biomolecular simulation program. J. Comput. Chem. 30, 1545–1614 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Tiwary, P. & Parrinello, M. A time-independent free energy estimator for metadynamics. J. Phys. Chem. B 119, 736–742 (2015).

    Article  CAS  PubMed  Google Scholar 

  85. Conicella, A.E., Zerze, G. H., Mittal, J. & Fawzi, N. L. ALS mutations disrupt phase separation mediated by α-helical structure in the TDP-43 low-complexity C-terminal domain. Structure 24, 1537–1549 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank M. Naik for helpful advice and assistance with NMR spectroscopy and V. Ryan for helpful discussions. We thank J. Ying for creating the HSQC-ROESY-HSQC experiment. Research was supported in part by NIGMS R01GM118530 (to N.L.F.), NIGMS R01GM120537 (to J.M.), Human Frontier Science Program RGP0045/2018 (to N.L.F.). A.C.M. was supported in part by NIGMS training grant to the MCB graduate program at Brown University (T32GM136566) and NSF graduate fellowship (1644760, to A.C.M.). Use of the high-performance computing capabilities of the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by the NSF grant TG-MCB-120014, is gratefully acknowledged. The content is solely the responsibility of the authors and does not necessarily represent the official views of the funding agencies.

Author information

Authors and Affiliations

Authors

Contributions

A.C.M. and N.L.F. designed, performed and analyzed data for NMR spectroscopy, phase-separation assays and microscopy. A.M.J. and D.H.S. performed fluorescence microscopy and provided reagents. T.M.P. assisted with reagents and recorded titration experiments. W.S.T., N.J. and J.M. designed and performed simulation experiments and analyzed the resulting data. A.C.M., J.M. and N.L.F. wrote the manuscript with comments from all authors.

Corresponding authors

Correspondence to Jeetain Mittal or Nicolas L. Fawzi.

Ethics declarations

Competing interests

N.L.F. is a member of the Scientific Advisory Board of Dewpoint Therapeutics LLC. A.C.M. is currently employed by Genentech. The authors declare no other competing interests.

Additional information

Peer review information Nature Structural & Molecular Biology thanks the anonymous reviewers for their contribution to the peer review of this work. Beth Moorefield was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

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

Extended data

Extended Data Fig. 1 RGG lysine variants are still capable of LLPS.

a) DIC micrographs of 5 μM MBP-FUS FL WT, RGG1 R9xK, RGG2 R8xK, and RGG3 R10xK after cleavage of the N-terminal MBP solubility tag by addition of TEV protease (left) or in the absence of TEV protease (right) in buffer containing 1 M sodium chloride. Scale bars are 20 μm. b) Turbidity over time of 5 μM MBP-FUS FL WT, RGG1 R9xK, RGG2 R8xK, and RGG3 R10xK after cleavage of the N-terminal MBP solubility tag by addition of TEV protease in buffer containing 1 M sodium chloride over time. The data are blanked to samples lacking TEV protease. Data are plotted as mean ± s.d of measurements from n = 3 replicates in one representative data set out of two independent experiments. c) DIC micrographs of 5 μM MBP-FUS FL WT, RGG1 R9xK, RGG2 R8xK, and RGG3 R10xK after cleavage of the N-terminal MBP solubility tag by addition of TEV protease (left) or in the absence of TEV protease (right) in the presence of 1:1 mass equivalents of total yeast RNA. Scale bars are 20 μm. d) Turbidity of 5 μM MBP-FUS FL WT, RGG1 R9xK, RGG2 R8xK, and RGG3 R10xK in the presence of 1:0, 1:1 and 1:2 total yeast RNA over time. The data are blanked to samples lacking TEV protease. Data are plotted as mean mean ± s.d. of measurements from n = 3 replicates in one representative data set out of two independent experiments.

Source data

Extended Data Fig. 2 The RGG domains weakly interact with the SYGQ LC.

a) 15N chemical shift perturbations and intensity differences of the SYGQ LC in the presence of increasing concentrations of MBP-RGG1, MBP-RGG2, MBP-RGG3 or MBP alone (negative control). Intensity data are normalized to a SYGQ LC alone control and are plotted as mean mean ± s.d. of baseline noise for each spectrum as estimate of uncertainty in one representative data set out of two independent experiments. b) 15N chemical shift perturbations and intensity differences of RGG1, RGG2, or RGG3 with increasing concentrations of MBP-SYGQ LC. The data are relative to RGG1, RGG2 or RGG3 alone controls. The asterisks for the RGG1 and RGG2 titrations with MBP indicate where the data are normalized to the 1:1 condition. Gray bars represent RGG motifs. Black dots correspond to resonances that are unassigned, while gray dots represent resonances that are assigned but not resolved due to overlap. Intensity data are plotted as mean ± s.d. of baseline noise for each spectrum as estimate of uncertainty in one representative data set. C) Average 15N chemical shift perturbations across all positions in SYGQ LC in the presence of ten times excess MBP-RGG1, MBP-RGG2, MBP-RGG3 or MBP alone (negative control) (full data points presented in a). Data are plotted as mean ± s.e.m. in one representative data set out of two independent experiments. d) Average 15N chemical shift perturbations across all positions in RGG1, RGG2 or RGG3 in the presence of ten times excess MBP-SYGQ LC or MBP alone (negative control) (full data points presented in b). Data are plotted as mean ± s.e.m. in one representative data set out of two independent experiments.

Source data

Extended Data Fig. 3 Assigned spectra of FUS RGG domains and impact of RGG mutations on weak interactions with SYGQ LC.

a-c) Assigned 1H-15N HSQC spectra of FUS RGG1, RGG2 or RGG3 in the dispersed phase. d) 15N chemical shift perturbations and intensity differences of 30 μM SYGQ LC in the presence of 300 μM of MBP-RGG3 WT, R10xK or R10xS. Intensity data are normalized to a SYGQ LC alone control and are plotted as mean ± s.d. of baseline noise for each spectrum as estimate of uncertainty in one representative data set. e) Average 15N chemical shift perturbations of SYGQ LC in the presence of ten times excess MBP-RGG3 WT, R10xK or R10xS (full data points presented in d). Data are plotted as mean ± s.e.m. in one representative data set. f) 15N transverse relaxation rate constant values for SYGQ LC in the presence of ten times excess of MBP (negative control), MBP-RGG1, MBP-RGG2 or MBP-RGG3. Data are plotted as mean ± propagated best-fit parameter confidence interval equal to 1 s.d in one representative data set.

Source data

Extended Data Fig. 4 Composition of fragments used for all-atom simulations.

a) Amino acid content of the SYGQ LC and fragments 11-54 and 120-163 used for all-atom simulations. b) Amino acid content of the FUS RGG domains. Fragments used for all-atom simulations (RGG1 220-267, RGG2 372-419 and RGG3 454-501) contain similar amino acid compositions to their experimental counterparts. c) Amino acid content of RNA polymerase C-terminal tail heptads 27-52. d) 15N spin relaxation parameters for FUS RGG3 in the dispersed phase from experiment and simulations. The segments used for simulations are shorter, explaining the discrepancies at the termini. Experimental data are plotted as mean ± propagated best-fit parameter confidence interval equal to 1 s.d in one representative data set of two independent experiments. Simulated data are plotted as mean ± s.e.m of n = 12 independent trajectories launched from randomly selected equilibrated ensemble members. E) Free energy landscape as a function of van der Waals contacts formed between hydrophobic atoms in FUS 11-54 or FUS 120-163 and RGG1, RGG2 or RGG3 from simulations. The use of different 44-amino acid fragments of FUS LC in the simulations produces differences in the energy landscapes, suggesting that the amino acid variation between the fragments used can have an impact on the number of contacts. Data are plotted as mean ± s.e.m of n = 5 equal divisions of the total data set from one data set with n = 16 independent replicas using PTWTE. F) Radius of gyration distribution of three different 44-residue long RGG fragments in single-chain simulations. The differences in compaction within the simulation system reflects the differences in amino acid composition of each RGG fragment. Data are plotted as mean ± s.e.m of n = 5 equal divisions of the total data set as in (e).

Source data

Extended Data Fig. 5 NOEs within a two-component condensed phase containing FUS SYGQ LC and RGG3.

a) NOE build-up curve (NOE intensity vs mixing time, τm) from 4D HSQC–NOESY-HSQC experiments. No diagonal peaks are present in these HSQC-NOESY-HSQC spectra, so data were collected as one-dimensional experiments and presented here as integration over the resonance envelope. Each experiment was performed once. b) 2D-planes from a 13C-HSQC-NOESY-15N-HSQC experiment recorded with a NOESY mixing time of 50 ms. c) Intermolecular ROEs from SYGQ LC are observed for arginine and other residue types including glycine in the 2-component condensed phase. 2D-projection from a 4D HSQC-NOESY-HSQC (250 ms mixing time; unscaled and scaled to match ROE) and HSQC-ROESY-HSQC (5 kHz spin lock / mixing for 20 ms). ROESY spin lock mixing time was limited due to more rapid transverse relaxation rate as compared to the NOESY mixing time, as the magnetization is longitudinal during the NOE transfer but transverse during the spin-locked ROE transfer. Experiments performed once. d) 1H-13C HSQC of FUS RGG3 in the dispersed phase. e) NOE signal intensity quantification from a 12C-filtered, 13C-edited NOESY-HSQC experiment presented in Fig. 3c. Intensity data for one representative experiment are plotted as mean ± s.d. of baseline noise for each plane as estimate of uncertainty in one representative data set.

Source data

Extended Data Fig. 6 Contacts between FUS SYGQ LC and RGG domains.

a) Total intermolecular contact propensities from two-chain simulations of SYGQ LC11-54 and RGG1220-267 binned by residue position (left), binned by residue type (center), and binned by residue type and normalized by residue frequency (right). Plots represent the total number of contacts for a particular residue position. Bars represent the total number of contacts for a particular residue type. Residues colored in gray occur in the sequence less than three times. (For A,B,C: Data are plotted as mean ± s.e.m of (left) n = 5 equal divisions of the total 16 replica PTWTE data set, (middle) total contact propensities, or (right) normalized total contact propensities from one representative data set out of two independent experiments.) b) Inter-residue contact propensities from two-chain simulations of SYGQ LC11-54 and RGG2372-419 binned by residue position (left), binned by residue type (center), and binned by residue type and normalized by residue frequency (right). Curved plots represent the total contact propensities for each residue. Bars represent the total number of contacts for a particular residue type. Gray bars represent residue types that occur less than three times in the sequence. c) Inter-residue contact propensities from two-chain simulations of SYGQ LC11-54 and RGG3454-501 binned by residue position. Plots represent the total contact propensities for each residue. Corresponding residue typed binned and frequency normalized plots (matching middle and right plots, respectively, for B and C) are presented in main text Fig. 3d,e. d) Total sp2/π interactions (left) and normalized by all VdW contacts (right) where all geometries are included (only distance-based definition) in two-chain simulations of SYGQ LC11-54 or SYGQ LC120-163 with RGG1, RGG2 or RGG3. The data are binned for π-π (top, lightest), sp2−π (middle, lighter) and sp2−sp2 (bottom) contacts. Data are plotted as mean ± s.e.m of n = 5 equal divisions of the total data set. e,f) Top fifteen interacting amino acid pairs in order of highest to lowest contact frequency (left to right) SYGQ LC11-54 or SYGQ LC120-163 with RGG1 or RGG2. The fraction of pairs showing hydrogen bonds, sp2/π, and cation-π contacts out of the total pairs with van der Waals interactions is indicated.

Source data

Extended Data Fig. 7 Contacts within a three-component phase containing FUS SYGQ LC and RGG3 and RNAP2 CTD.

a) Chemical shift perturbations and signal intensity changes for 15N-RNA polymerase II CTD in the presence of increasing concentrations of FUS RGG3. Intensity data are plotted as mean ± s.d. of baseline noise for each spectrum as estimate of uncertainty in one representative data set. b) Free energy landscape of van der Waals contacts between hydrophobic atoms between RNAP2 CTD and FUS SYGQ LC or RGG3 from two-chain simulations. Data are plotted as mean ± s.e.m of n = 5 equal divisions of the total data set from one representative data set. c) 1H-15N HSQC of RNA polymerase II CTD in the dispersed (orange) and condensed (green) phases. d) NOE signal intensity quantification from a 12C-filtered, 13C-edited NOESY-HSQC experiment presented in Fig. 6b. Intensity data are plotted as mean ± s.d. of baseline noise for each plane as estimate of uncertainty in one representative data set out of two independent experiments. Inter-residue contact propensities from two-chain simulations of RNAP2 CTD1853-1896 and e) FUS SYGQ LC11-54 or f) RGG3 binned by residue position. Plots represent the total number of contacts for a particular residue position. Data are plotted as mean ± s.e.m of n = 5 equal divisions of the total data set from one representative data set.

Source data

Supplementary information

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

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 Fig. 7

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Murthy, A.C., Tang, W.S., Jovic, N. et al. Molecular interactions contributing to FUS SYGQ LC-RGG phase separation and co-partitioning with RNA polymerase II heptads. Nat Struct Mol Biol 28, 923–935 (2021). https://doi.org/10.1038/s41594-021-00677-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-021-00677-4

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