Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2018 Sep 14;115(40):9929–9934. doi: 10.1073/pnas.1804177115

Relation between single-molecule properties and phase behavior of intrinsically disordered proteins

Gregory L Dignon a,1, Wenwei Zheng b,c,1, Robert B Best c, Young C Kim d, Jeetain Mittal a,2
PMCID: PMC6176625  PMID: 30217894

Significance

Some proteins can undergo phase separation in vivo to form membraneless organelles with distinct physiological functions. In many cases, the driving force is believed to be the intrinsically disordered domains that lack regular structural features and can phase separate in vitro. Relating protein sequences to phase-separation propensity is the goal of this work. Toward this goal, we highlight the connection between the temperature where a single chain collapses, the temperature where interactions within a dilute solution favor self-association, and the critical temperature of phase separation. Our results show that these characteristic temperatures are highly correlated, suggesting that experiments performed in dilute conditions may be used to predict phase separation.

Keywords: membraneless organelles, liquid–liquid phase separation, intrinsically disordered proteins

Abstract

Proteins that undergo liquid–liquid phase separation (LLPS) have been shown to play a critical role in many physiological functions through formation of condensed liquid-like assemblies that function as membraneless organelles within biological systems. To understand how different proteins may contribute differently to these assemblies and their functions, it is important to understand the molecular driving forces of phase separation and characterize their phase boundaries and material properties. Experimental studies have shown that intrinsically disordered regions of these proteins are a major driving force, as many of them undergo LLPS in isolation. Previous work on polymer solution phase behavior suggests a potential correspondence between intramolecular and intermolecular interactions that can be leveraged to discover relationships between single-molecule properties and phase boundaries. Here, we take advantage of a recently developed coarse-grained framework to calculate the θ temperature Tθ, the Boyle temperature TB, and the critical temperature Tc for 20 diverse protein sequences, and we show that these three properties are highly correlated. We also highlight that these correlations are not specific to our model or simulation methodology by comparing between different pairwise potentials and with data from other work. We, therefore, suggest that smaller simulations or experiments to determine Tθ or TB can provide useful insights into the corresponding phase behavior.


Liquid–liquid phase separation (LLPS) of proteins and nucleic acids has recently drawn significant attention due to its relevance to physiological functions (14), disease pathology (57), and design of self-assembling materials with tunable properties (810). One major reason for this interest in LLPS is its relation to intracellular compartmentalization via the formation of membraneless organelles (1114). Experimental studies suggest that the main driving force of LLPS for many of the proteins involved comes from their intrinsically disordered domains (i.e., those that lack a stable folded structure) (11, 12). A working hypothesis is that the disordered domains drive LLPS under physiological conditions, thus increasing the effective concentration of other (folded) domains (e.g., RNA binding domains) that carry out additional functions by recruiting other biomolecules, such as RNA (11, 1417). Many proteins that are known to form dynamic liquid-like assemblies through LLPS can also form pathological aggregates. It has been suggested that this high-density phase might help overcome nucleation barriers and promote aggregation (6, 7). The effects of known disease-related mutations, which promote both LLPS and formation of solid-like aggregates, provide evidence in support of this hypothesis (7, 17). LLPS-susceptible protein sequences are also promising for the design of materials for a variety of applications (8, 9).

Relating sequence to phase-separation properties is a major goal in this work; it will allow for the exploration of the proteome to identify sequences which may participate in LLPS and will aid in directed design of peptide sequences having desired assembly and material properties. An excellent study by Quiroz and Chilkoti (8) has already provided much evidence on the relationship between amino acid composition and phase behavior by measuring the demixing temperature for a large number of sequences. Several studies have also addressed the relation between the degree of collapse of an intrinsically disordered protein (IDP) in dilute conditions and the phase boundaries. Lin and coworkers (18, 19) observed a correlation between the radius of gyration (Rg) of an isolated protein and the LLPS critical temperature (Tc) for a collection of polyampholytic sequences. Riback et al. (20) found a similar correlation between the single-chain Rg and the demixing temperature (Tdemix) for sequences of differing hydrophobicity. Both of these studies, despite focusing on different aspects of protein sequence properties, suggest that the Rg of a single chain in dilute solution is linked to its ability to phase separate. However, such a correspondence is not expected to hold well for proteins of different lengths due to the differences in the scaling of Rg and Tc as a function of chain length.

LLPS of IDPs shares common features with the well-established field of polymer solution phase behavior, suggesting that existing polymer-physics principles might be a good starting point toward a better understanding of the underlying phenomenon (8, 2123). For example, the collapse of a single isolated homopolymer chain and polymer solution phase separation are related via the effective monomer–monomer interaction strengths in the Flory–Huggins theory (24, 25). Panagiotopoulos et al. (26) have shown that the temperature of the coil-to-globule transition (Tθ) and Tc are equal in the limit of infinite chain length. Wang and Wang (27) applied self-consistent field theory to also show the correspondence between Tθ, the Boyle temperature (TB) at which the second virial coefficient (B22), is equal to 0, and Tc. We investigate whether these theories in polymer physics can also be applied to understand LLPS of finite-length heteropolymeric IDPs.

Our goal is to find a general relationship between Tθ, TB, and Tc for a variety of IDP sequences. To accomplish this, we use a recently developed coarse-grained (CG) computational framework that is capable of obtaining all three of these properties (28). We have shown previously that this approach is sufficient to capture qualitative trends in phase behavior as induced by sequence mutations or by inclusion of a folded domain (2830). We calculate these three characteristic temperatures for 20 different IDP sequences and mutants with diverse hydropathy, charge, patterning, and sequence length from experimental and theoretical studies on IDP phase separation (12, 17, 28, 30, 31). We further calculate Tθ for a set of 30 synthetic polyampholyte sequences and compare with Tc determined from theoretical methods to show that this correlation is not specific to our CG framework (18). We observe a strong linear correlation among Tθ, TB, and Tc. This highlights the utility of polymer-physics principles in discovering predictive models of protein LLPS. Moreover, since it may be easier to obtain Tθ and TB via simulations or some experimental methods, they can serve as proxy descriptors of LLPS and allow for high-throughput screening of sequences.

Results and Discussion

Obtaining Phase Diagrams from Molecular Simulations.

To obtain the phase coexistence envelope for an IDP sequence, we use a recently proposed CG modeling framework (28) that has already been applied to understand the sequence determinants of specific IDPs (29, 30). The sequences that we consider include the low-complexity (LC) domain of FUS and four different phosphomimetic variants (7, 12, 29), the multivalent repetitive FUS sequences ([FUS40]n) (28), the LC domain of hnRNPA2 and two disease-related mutants (30, 32, 33), the N-terminal intrinsically disordered region of LAF-1 (11) and five designed variants having similar or identical sequence composition, and the disordered C-terminal domain of TDP-43 (17). We obtain the phase coexistence using two different CG potentials for several of the sequences to check whether the results presented here are model-dependent.

For efficient sampling of the phase diagram, we use the slab method (28, 34), in which the simulations are initiated from a high-density protein slab that is continuous in the x,y plane of a simulation cell with an elongated z axis, allowing for efficient equilibration between the low- and high-density phases (Fig. 1C and Movie S1). This geometry has been widely used to study liquid–liquid (35), solid–liquid (36), and liquid–vapor (37) phase coexistence, and the results are highly consistent with conventional grand canonical approaches (34, 38) (SI Appendix, Fig. S1). Standard chemical potential routes are computationally prohibitive for longer chains that we study here. This challenge can be overcome by the use of lattice models (26) or by simplifying the protein chains as patchy particles as done recently by Nguemaha and Zhou (39). We instead prefer an approach that can faithfully capture the polymeric nature of IDPs and their interactions leading to the formation of a dense phase with concentration that is dependent on the protein sequence.

Fig. 1.

Fig. 1.

Determining phase diagrams via slab simulations. (A) Density profiles become flat at temperatures higher than Tc. Inset shows low-density phase in log scale to highlight that the density converges well and is nonzero. (B) From the density profile, we generate the coexistence curve of the sequence. (C) Snapshots of slab simulations at supercritical (Top), near-critical (Middle), and below-critical (Bottom) conditions.

We show in SI Appendix, Fig. S2 that simulations of phase coexistence using a cubic box with a spherical “droplet” of chains (Movie S2) and using slab geometry produce very similar phase diagrams. By simulating the droplet geometry at different system sizes, we also find that the coexistence densities approach the values determined using slab geometry as the system size increases, even past the number of chains used for the slab simulations themselves (SI Appendix, Fig. S3). We conclude that the slab geometry provides an advantage over simulations using droplet geometry in that it reduces finite-size effects.

For WT hnRNPA2 at temperatures where we observe phase coexistence, we calculate concentrations in the low-density phase to be in the range of 0.017–23.8 mg/mL (1.2–1,870 μM) and the high-density phase to be within 324.0–664.3 mg/mL (22.6–43.4 mM). These densities are in excellent agreement with experimental measurements by Ryan et al. (30), who observed micromolar concentration of the aqueous phase at low temperatures and estimated the concentration in the condensed phase to be between 30 and 40 mM. Brady et al. (14) also find the coexistence concentrations to be within the ranges of 0–50 and 150–400 mg/mL for the low- and high-density phases, respectively, for Ddx4, a sequence similar to LAF-1. We see good agreement of coexistence densities with many other experimental studies as well that show proteins having saturation concentrations in the 1-mg/mL range and high-density concentration in the hundreds of milligrams per milliliter range (12, 17, 29, 40).

As the nonbonded interactions between amino acid residues in our models are temperature independent, we only observe the upper critical solution temperature, hereafter referred to as Tc. We calculate Tc from the coexistence densities as detailed in SI Appendix. We note that the Tc estimated in this way is rather insensitive to the range of temperatures fitted (SI Appendix, Fig. S4). We have also directly verified Tc estimated by this method in several cases by running simulations near Tc: just 5 K above Tc, the slab is observed to disintegrate, while below Tc, it is stable and will spontaneously form even when the simulation is initiated from a fully dispersed initial condition (SI Appendix, Fig. S5 and Movie S3). Representative coexistence density profiles and the phase diagram are shown in Fig. 1 A and B for WT hnRNPA2 along with snapshots of simulations at different temperatures (Fig. 1C). Phase diagrams for all sequences are presented in SI Appendix, Fig. S6.

The Relation Between Single-Chain Properties and Protein LLPS.

We further determine Rg as a function of temperature from replica exchange molecular dynamics (REMD) simulations of a single CG protein chain (Movie S4). We note that the models used here provide chain dimensions and asphericity values similar to an all-atom implicit solvent model (41), which has been applied extensively to IDPs involved in LLPS (4245) (SI Appendix, Fig. S7). We also note that, using the Kim–Hummer (KH) variant of our coarse-grained potential, the Rg values of the two sequences, hnRNPA2 and FUS, are 2.2 and 3.4 nm, respectively, at 300 K, which are comparable with experimental measurements of the hydrodynamic radius by Ryan et al. (30) of 2.89 and 3.32 nm, respectively. When comparing WT and phosphomimetic variants of FUS, we see the expected increase of chain dimensions as also reported by Ryan et al. (30). This increase in chain dimensions being accompanied by lower propensity for phase separation has been observed by several other studies (18, 20, 29) and should hold for sequences of similar length and composition.

For IDPs of different lengths, Rg scales with respect to chain length as RgNν, where N is the number of peptide bonds and ν is the Flory scaling exponent (46). Since Rg by itself is not expected to correlate well with Tc for IDPs of different length (SI Appendix, Fig. S8), we use ν, which we believe should scale more similarly to Tc with changes in chain length. ν can be determined directly from single-chain simulations by fitting average interresidue distance Rij between the ith and jth residues as a function of chain separation |ij| (47) as follows, Rij=b|ij|ν, where b is the Kuhn length, which is set to be 0.55 nm as suggested for disordered proteins (4850) (Fig. 2A). It should be noted that, for collapsed proteins, Rij can deviate from the power law scaling (low-temperature data in Fig. 2A) due to intramolecular interactions resulting in the formation of nonspecific structures. We observe good fits for data near and above Tθ for most sequences as shown in Fig. 2A. For each protein, ν is estimated as a function of the temperature to determine Tθ as the temperature where ν=1/2. An example is shown in Fig. 2B for WT hnRNPA2 (32, 33); results for all other proteins are given in SI Appendix, Fig. S9. At Tθ, the protein chain behaves like an ideal chain, as attractive intraprotein interactions are canceled by repulsive excluded volume interactions (46).

Fig. 2.

Fig. 2.

Comparison between Tθ obtained from single-chain simulations and Tc obtained from slab simulations. (A) Scaling of average intramolecular distances, Rij, with sequence separation |ij|, as a function of temperature. The dashed green, black, and blue curves correspond to scaling exponents of 1/3, 1/2, and 3/5, respectively. Snapshots in Insets show an expanded conformation under good solvent conditions and a collapsed configuration under poor solvent conditions. (B) ν crosses 0.5 at the Θ point and defines Tθ. (C) Correlation between Tc and Tθ. The error bars of both quantities are smaller than the symbols.

Previous studies have shown that Tθ of a homopolymer coincides with Tc in the limit of infinite chain length (26, 27). We are interested in determining whether finite-length chains of heteropolymers, such as IDPs, may still show such a correspondence between the two temperatures. We test if solvent quality characterized by ν at a single temperature (300 K) can provide useful information about Tc. The overall correlation between Tc and ν (SI Appendix, Fig. S10) is quite reasonable (R2=0.760). We observe an even stronger correlation (R2 = 0.925) between Tθ and Tc as shown in Fig. 2C for all sequences with temperatures varying over a broad range from 200 to 500 K. The correlation between Tθ and Tc is found to be approximately linear with a slope of one and intercept near the origin, which suggests that the temperature at which intramolecular attraction and repulsion are balanced is roughly the same as the temperature at which intermolecular interactions in a protein solution are also canceled out. Thus, we expect that this correlation should not change significantly for experimental data on similar protein sequences and hope that our results will motivate future experimental work in this direction. It is also notable that the correlation is independent of the computational model used, as we include data from two very different CG models of interresidue interactions (Figs. 2C and 4C and SI Appendix, Fig. S11), further suggesting that these correlations could be universal, regardless of technique or protein.

Fig. 4.

Fig. 4.

Comparison between Boyle temperature TB obtained from the temperature dependence of B22 and Tc. (A) Radial distribution function between the two isolated chains with increasing temperature. The snapshot in Inset shows typical configurations in the simulation. (B) The temperature at which B22=0 defines TB. (C) Correlation between Tc and TB. The errors of TB are shown as bars, whereas the errors of Tc are smaller than the symbols.

The protein sequences considered so far do not account for additional factors that are important for the IDP behavior, such as charge patterning (47). To check if the correlation between Tc and Tθ can also be useful for capturing changes in the protein LLPS due to sequence patterning, we consider a set of 30 polyampholytic sequences with the same amino acid composition but differently arranged, ranging from perfectly mixed to block copolymer (51, 52). Lin and coworkers (18, 22) recently investigated the LLPS behavior of these protein sequences (termed SV series) via Flory–Huggins theory combined with a random phase approximation method. We take advantage of the available Tc data from this study by comparing with Tθ from our model using REMD simulations. We find the correlation to be quite strong (R2 = 0.974) for the SV series as shown in Fig. 3. We note that, for sequences having a high degree of charge segregation, the intrachain distances do not fit the polymer scaling law well (SI Appendix, Fig. S12). Therefore, we have also applied an alternative method to obtain ν (50) (SI Appendix) and shown that Tθ, when calculated by this method, also correlates very well with the Tc results from Lin and coworkers (18, 22) (SI Appendix, Fig. S13). The correlation is better than those seen for the previous 20 IDP sequences, mainly due to the balanced charge and identical sequence composition. These results suggest that well-established polymer-physics principles can be used to understand and predict the LLPS of IDPs. In general, when considering intramolecular interactions, as in the case of Tθ, each individual residue’s self-interactions are ignored. Such interactions may be quite relevant for intermolecular interactions when estimating Tc, especially for sequences that are not well-mixed or charged sequences with long-range electrostatic interactions. In the case of LAF-1, we had previously identified a short eight-residue region that has high propensity for interactions with itself in other chains. Moreover, LAF-1 also has a higher fraction of charged residues that will further make it difficult to capture information on such interactions purely based on a single-chain property, such as Tθ. We expect that such interactions will be better captured with measures, such as second virial coefficient.

Fig. 3.

Fig. 3.

Test on designed sequences with different charge patterns (51). Tc, as calculated by Lin and Chan (18), and Tθ, as determined from our monomer simulations and normalized to the sequence with the lowest temperature (sv1), are plotted against each other and show good correlation.

The Relation Between Protein Affinities in Dilute Solutions and LLPS.

The observed correlation between Tθ and Tc suggests a strong similarity between the intrachain interactions of an isolated protein and the interchain interactions among proteins in a liquid-like phase. Interchain interactions can also be quantified by the osmotic second virial coefficient (B22). Previously, B22 was shown to be related to the growth rates of protein crystals (53) and the temperature at which protein solutions become turbid (54). As discussed in the Introduction, Tθ of a homopolymer chain coincides with TB in the limit of infinite chain length (26, 27). Therefore, as in the previous section, we would like to test if such a correlation applies to finite-length IDPs that are involved in LLPS.

In molecular simulations, B22 can be calculated from the radial distribution function, g(r), between two protein chains (55) as follows:

B22=2π0[1g(r)]r2dr. [1]

It is somewhat cumbersome to estimate B22 from simulation for disordered proteins, as the g(r) sampling involves conformational changes as well as distance and orientational degrees of freedom. However, it is still considerably less computationally demanding than sampling phase coexistence using slab geometry. Following previous work (56), we are able to determine B22 with excellent statistical certainty (especially near the TB) using a combination of Monte Carlo and umbrella sampling methods (Movie S5). The g(r) for WT hnRNPA2 is shown in Fig. 4A. We determine TB by interpolating the temperature at which B22=0 in Fig. 4B for WT hnRNPA2 and in SI Appendix, Fig. S14 for all of the proteins (details are in SI Appendix,). We observe a strong correlation between TB and Tc (R2 = 0.965) as shown in Fig. 4C (the correlation between TB and Tθ is shown in SI Appendix, Fig. S15). The correlation between Tc and TB seems to be statistically stronger than between Tc and Tθ, which suggests that, for finite-length proteins, intermolecular interactions captured by B22 may be a better metric than ν that is based on intrachain interactions. Of course, the two temperatures differ from each other for the finite-length proteins considered here compared with the infinitely long homopolymers. The slope of the correlation is not one and its intercept is not zero as with Tc and Tθ. Considering that the results from different models fall on the same correlation, it would be tempting to conclude that the slope and intercept of the fit to the data will remain unchanged if experimental data were also included here. However, given the lack of suitable theories that can be used to assess this issue further, we reserve such judgment for future work.

Conclusions

We have shown in this work that theories originally intended for use with homopolymers of infinite chain length (26, 57) can also be applied to heteropolymeric, finite-length IDPs to relate the characteristics of dilute and condensed phase interactions. Specifically, we have observed strong correlations between Tc, Tθ, and TB. This general correlation is encouraging for several reasons. (i) The correlation suggests that our knowledge of homopolymers and the variety of tools from polymer physics are applicable to the study of LLPS of IDPs. (ii) The correlation can advance the simulation prediction of Tc by using either more detailed or faster models. All-atom explicit solvent simulations, which predict single-chain dimensions with high accuracy (5860) but are also more computationally demanding, can be applied to the study of phase behavior by taking advantage of the correlation between Tc and Tθ. Rapid screening of IDP sequences for their phase behavior might be also possible if Tθ can be predicted by sequence composition (e.g., mean hydrophobicity or net charge) or by efficient CG simulations. More work is clearly needed to reach this ambitious goal. (iii) This relation allows a variety of alternative experimental techniques, such as single-molecule FRET (61, 62) and small-angle X-ray scattering (SAXS) (20, 63), to be used to infer the relative propensity of different protein sequences to undergo LLPS. It is possible to estimate ν using single-molecule FRET, as previously done by Schuler and coworkers (48) with multiple FRET dye labeling positions. In addition, recently developed methods for analysis of SAXS (50, 64) and FRET (65) allow a value of ν to be estimated more easily from a single variant of any protein with the same experimental burden as would be needed to obtain Rg. In SI Appendix, Figs. S10 and S11, we present correlations of Tc with alternative experimental observables that may be easier to determine, depending on the context.

One can, in principle, also extend this beyond critical temperature and estimate critical salt concentration, critical pH, and other solution conditions just from the response of ν and B22 to perturbations. An additional advantage of using TB as an indicator of phase separation is that it has the potential to be extended to characterize coassembly of mixtures of different proteins and of protein and RNA mixtures, which is of great biological interest (13, 66). However, there are going to be examples where interactions are more complicated than what can be captured at the level of one or two chains using Tθ or TB, such as systems driven by higher-order oligomerization of folded domains (13, 67). Despite these caveats, we trust that our results here will promote rigorous characterization of IDPs at different conditions and allow the field to progress toward solving the problem of sequence-encoded phase behavior. Such a high-throughput method would allow for rapid design of sequences, such as those designed by Quiroz and Chilkoti (8), and would aid in the development of protein models and techniques that can be used to more accurately predict the properties of IDP LLPS (13, 28, 68).

Materials and Methods

CG Model.

We use our recently developed Cα-based model, where proteins are represented as flexible chains and each amino acid residue is considered as a single particle. Bonds are modeled using harmonic springs, long-range electrostatics are modeled using a Coulombic term with Debye–Hückel electrostatic screening (69), and nonbonded pairwise interactions are modeled using one of two knowledge-based potentials that we have previously applied to these systems (28, 56, 70) (SI Appendix).

Simulation Methods.

An initial configuration for slab simulations was generated via simulations at constant pressure and temperature, starting from a dispersed phase of protein chains and collapsed to a dense phase. Production simulations were conducted for 5 μs at constant temperature and volume using the HOOMD-Blue v2.1.5 software package (71). Monomer simulations were conducted at a range of temperatures using REMD (72) in cubic boxes using the LAMMPS software package (73). Umbrella-sampling Monte Carlo simulations were used for simulations of two protein chains using in-house simulation software. Errors for Tc and Tθ are estimated using block averages with five blocks, and for TB, we follow a bootstrapping strategy. Additional details are in SI Appendix.

Supplementary Material

Supplementary File
Supplementary File
Supplementary File
pnas.1804177115.sm02.gif (30.7MB, gif)
Supplementary File
pnas.1804177115.sm03.gif (12.8MB, gif)
Supplementary File
Supplementary File

Acknowledgments

This research is supported by US Department of Energy, Office of Science, Basic Energy Sciences Award DE-SC0013979. The data on FUS proteins were generated as part of a project supported by National Institutes of Health (NIGMS R01GM118530). This research used resources of the National Energy Research Scientific Computing Center (a Department of Energy Office of Science User Facility supported under Contract DE-AC02-05CH11231), the Extreme Science and Engineering Discovery Environment (XSEDE; supported by National Science Foundation Project TG-MCB120014), and the NIH High-Performance Computing Biowulf cluster (hpc.nih.gov). W.Z. acknowledges startup support from Arizona State University. R.B.B. was supported by the Intramural Research Program of the National Institute of Diabetes and Digestive and Kidney Diseases of the NIH. Y.C.K was supported by the Office of Naval Research via the US Naval Research Laboratory Base Program.

Footnotes

The authors declare no conflict of interest.

This article is a PNAS Direct Submission. H.S.C. is a guest editor invited by the Editorial Board.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1804177115/-/DCSupplemental.

References

  • 1.Brangwynne CP, et al. Germline p granules are liquid droplets that localize by controlled dissolution/condensation. Science. 2009;324:1729–1732. doi: 10.1126/science.1172046. [DOI] [PubMed] [Google Scholar]
  • 2.Li P, et al. Phase transitions in the assembly of multi-valent signaling proteins. Nature. 2012;483:336–340. doi: 10.1038/nature10879. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Su X, et al. Phase separation of signaling molecules promotes t cell receptor signal transduction. Science. 2016;352:595–599. doi: 10.1126/science.aad9964. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Hnisz D, Shrinivas K, Young RA, Chakraborty AK, Sharp PA. A phase separation model for transcriptional control. Cell. 2017;169:13–23. doi: 10.1016/j.cell.2017.02.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Xiang S, et al. The lc domain of hnRNPA2 adopts similar conformations in hydrogel polymers, liquid-like droplets, and nuclei. Cell. 2015;163:829–839. doi: 10.1016/j.cell.2015.10.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Molliex A, et al. Phase separation by low complexity domains promotes stress granule assembly and drives pathological fibrillization. Cell. 2015;163:123–133. doi: 10.1016/j.cell.2015.09.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Patel A, et al. A liquid-to-solid phase transition of the als protein FUS accelerated by disease mutation. Cell. 2015;162:1066–1077. doi: 10.1016/j.cell.2015.07.047. [DOI] [PubMed] [Google Scholar]
  • 8.Quiroz FG, Chilkoti A. Sequence heuristics to encode phase behaviour in intrinsically disordered protein polymers. Nat Mater. 2015;14:1164–1171. doi: 10.1038/nmat4418. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Lau HK, Li L, Jurusik AK, Sabanayagam CR, Kiick KL. Aqueous liquid–liquid phase separation of resilin-like polypeptide/polyethylene glycol solutions for the formation of microstructured hydrogels. ACS Biomater Sci Eng. 2016;3:757–766. doi: 10.1021/acsbiomaterials.6b00076. [DOI] [PubMed] [Google Scholar]
  • 10.Simon JR, Carroll NJ, Rubinstein M, Chilkoti A, López GP. Programming molecular self-assembly of intrinsically disordered proteins containing sequences of low complexity. Nat Chem. 2017;9:509–515. doi: 10.1038/nchem.2715. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Elbaum-Garfinkle S, et al. The disordered p granule protein laf-1 drives phase separation into droplets with tunable viscosity and dynamics. Proc Natl Acad Sci USA. 2015;112:7189–7194. doi: 10.1073/pnas.1504822112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Burke KA, Janke AM, Rhine CL, Fawzi NL. Residue-by-residue view of in vitro FUS granules that bind the c-terminal domain of rna polymerase ii. Mol Cell. 2015;60:231–241. doi: 10.1016/j.molcel.2015.09.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Feric M, et al. Coexisting liquid phases underlie nucleolar subcompartments. Cell. 2016;165:1686–1697. doi: 10.1016/j.cell.2016.04.047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Brady JP, 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. 2017;114:E8194–E8203. doi: 10.1073/pnas.1706197114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Zhang H, et al. Rna controls polyq protein phase transitions. Mol Cell. 2015;60:220–230. doi: 10.1016/j.molcel.2015.09.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Nott TJ, Craggs TD, Baldwin AJ. Membraneless organelles can melt nucleic acid duplexes and act as biomolecular filters. Nat Chem. 2016;8:569–575. doi: 10.1038/nchem.2519. [DOI] [PubMed] [Google Scholar]
  • 17.Conicella AE, Zerze GH, Mittal J, Fawzi NL. Als mutations disrupt phase separation mediated by α-helical structure in the tdp-43 low-complexity c-terminal domain. Structure. 2016;24:1537–1549. doi: 10.1016/j.str.2016.07.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Lin YH, Chan HS. Phase separation and single-chain compactness of charged disordered proteins are strongly correlated. Biophys J. 2017;112:2043–2046. doi: 10.1016/j.bpj.2017.04.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Lin YH, Brady JP, Forman-Kay JD, Chan HS. Charge pattern matching as a‘fuzzy’mode of molecular recognition for the functional phase separations of intrinsically disordered proteins. New J Phys. 2017;19:115003. [Google Scholar]
  • 20.Riback JA, et al. Stress-triggered phase separation is an adaptive, evolutionarily tuned response. Cell. 2017;168:1028–1040. doi: 10.1016/j.cell.2017.02.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Brangwynne CP, Tompa P, Pappu RV. Polymer physics of intracellular phase transitions. Nat Phys. 2015;11:899–904. [Google Scholar]
  • 22.Lin YH, Forman-Kay JD, Chan HS. Sequence-specific polyampholyte phase separation in membraneless organelles. Phys Rev Lett. 2016;117:178101. doi: 10.1103/PhysRevLett.117.178101. [DOI] [PubMed] [Google Scholar]
  • 23.Lin YH, Song J, Forman-Kay JD, Chan HS. Random-phase-approximation theory for sequence-dependent, biologically functional liquid-liquid phase separation of intrinsically disordered proteins. J Mol Liq. 2017;228:176–193. [Google Scholar]
  • 24.Flory PJ. Thermodynamics of high polymer solutions. J Chem Phys. 1942;10:51–61. [Google Scholar]
  • 25.Huggins ML. Some properties of solutions of long-chain compounds. J Phys Chem. 1942;46:151–158. [Google Scholar]
  • 26.Panagiotopoulos AZ, Wong V, Floriano MA. Phase equilibria of lattice polymers from histogram reweighting Monte Carlo simulations. Macromolecules. 1998;31:912–918. [Google Scholar]
  • 27.Wang R, Wang ZG. Theory of polymer chains in poor solvent: Single-chain structure, solution thermodynamics, and θ point. Macromolecules. 2014;47:4094–4102. [Google Scholar]
  • 28.Dignon GL, Zheng W, Kim YC, Best RB, Mittal J. Sequence determinants of protein phase behavior from a coarse-grained model. PLoS Comput Biol. 2018;14:e1005941. doi: 10.1371/journal.pcbi.1005941. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Monahan Z, et al. Phosphorylation of FUS low-complexity domain disrupts phase separation, aggregation, and toxicity. EMBO J. 2017;36:2951–2967. doi: 10.15252/embj.201696394. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Ryan VH, et al. Mechanistic view of hnRNPA2 low-complexity domain structure, interactions, and phase separation altered by mutation and arginine methylation. Mol Cell. 2018;69:465–479.e7. doi: 10.1016/j.molcel.2017.12.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Wei MT, et al. Phase behaviour of disordered proteins underlying low density and high permeability of liquid organelles. Nat Chem. 2017;9:1118–1125. doi: 10.1038/nchem.2803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Kim HJ, et al. Mutations in prion-like domains in hnRNPA2b1 and hnRNPA1 cause multisystem proteinopathy and als. Nature. 2013;495:467–473. doi: 10.1038/nature11922. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Qi X, et al. Familial early-onset paget’s disease of bone associated with a novel hnRNPA2b1 mutation. Calcified Tissue Int. 2017;101:159–169. doi: 10.1007/s00223-017-0269-0. [DOI] [PubMed] [Google Scholar]
  • 34.Silmore KS, Howard MP, Panagiotopoulos AZ. Vapour–liquid phase equilibrium and surface tension of fully flexible lennard–jones chains. Mol Phys. 2017;115:320–327. [Google Scholar]
  • 35.Lynden-Bell R, Kohanoff J, Del Popolo M. Simulation of interfaces between room temperature ionic liquids and other liquids. Faraday Discuss. 2005;129:57–67. doi: 10.1039/b405514d. [DOI] [PubMed] [Google Scholar]
  • 36.Vega C, De Miguel E. Surface tension of the most popular models of water by using the test-area simulation method. J Chem Phys. 2007;126:154707. doi: 10.1063/1.2715577. [DOI] [PubMed] [Google Scholar]
  • 37.Müller EA, Mejía A. Comparison of united-atom potentials for the simulation of vapor–liquid equilibria and interfacial properties of long-chain n-alkanes up to n-c100. J Phys Chem B. 2011;115:12822–12834. doi: 10.1021/jp203236q. [DOI] [PubMed] [Google Scholar]
  • 38.Sheng YJ, Panagiotopoulos AZ, Kumar SK, Szleifer I. Monte Carlo calculation of phase equilibria for a bead-spring polymeric model. Macromolecules. 1994;27:400–406. [Google Scholar]
  • 39.Nguemaha V, Zhou HX. Liquid-liquid phase separation of patchy particles illuminates diverse effects of regulatory components on protein droplet formation. Sci Rep. 2018;8:6728 doi: 10.1038/s41598-018-25132-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Li HR, Chiang WC, Chou PC, Wang WJ, Huang Jr. Tar dna-binding protein 43 (tdp-43) liquid–liquid phase separation is mediated by just a few aromatic residues. J Biol Chem. 2018;293:6090–6098. doi: 10.1074/jbc.AC117.001037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Vitalis A, Pappu RV. Absinth: A new continuum solvent model for simulations of polypeptides in aqueous solutions. J Comput Chem. 2008;30:673–699. doi: 10.1002/jcc.21005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Pak CW, et al. Sequence determinants of intracellular phase separation by complex coacervation of a disordered protein. Mol Cell. 2016;63:72–85. doi: 10.1016/j.molcel.2016.05.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Harmon TS, Holehouse AS, Rosen MK, Pappu RV. Intrinsically disordered linkers determine the interplay between phase separation and gelation in multivalent proteins. Elife. 2017;6:e30294 doi: 10.7554/eLife.30294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Warner JB, IV, et al. Monomeric huntingtin exon 1 has similar overall structural features for wild-type and pathological polyglutamine lengths. J Am Chem Soc. 2017;139:14456–14469. doi: 10.1021/jacs.7b06659. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Posey AE, et al. Profilin reduces aggregation and phase separation of huntingtin N-terminal fragments by preferentially binding to soluble monomers and oligomers. J Biol Chem. 2018;293:3734–3746. doi: 10.1074/jbc.RA117.000357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Flory PJ. The configuration of real polymer chains. J Chem Phys. 1949;17:303–310. [Google Scholar]
  • 47.Mao AH, Crick SL, Vitalis A, Chicoine C, Pappu RV. Net charge per residue modulates conformational ensembles of intrinsically disordered proteins. Proc Natl Acad Sci USA. 2010;107:8183–8188. doi: 10.1073/pnas.0911107107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Hofmann H, et al. Polymer scaling laws of unfolded and intrinsically disordered proteins quantified with single-molecule spectroscopy. Proc Natl Acad Sci USA. 2012;109:16155–16160. doi: 10.1073/pnas.1207719109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Zheng W, et al. Probing the action of chemical denaturant on an intrinsically disordered protein by simulation and experiment. J Am Chem Soc. 2016;138:11702–11713. doi: 10.1021/jacs.6b05443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Zheng W, et al. Inferring properties of disordered chains from fret transfer efficiencies. J Chem Phys. 2018;148:123329. doi: 10.1063/1.5006954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Das RK, Pappu RV. Conformations of intrinsically disordered proteins are influenced by linear sequence distributions of oppositely charged residues. Proc Natl Acad Sci USA. 2013;110:13392–13397. doi: 10.1073/pnas.1304749110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Sawle L, Ghosh K. A theoretical method to compute sequence dependent configurational properties in charged polymers and proteins. J Chem Phys. 2015;143:085101. doi: 10.1063/1.4929391. [DOI] [PubMed] [Google Scholar]
  • 53.George A, et al. [6] Second virial coefficient as predictor in protein crystal growth. Methods Enzymol. 1997;276:100–110. doi: 10.1016/S0076-6879(97)76052-X. [DOI] [PubMed] [Google Scholar]
  • 54.Platten F, Hansen J, Wagner D, Egelhaaf SU. Second virial coefficient as determined from protein phase behavior. J Phys Chem Lett. 2016;7:4008–4014. doi: 10.1021/acs.jpclett.6b01714. [DOI] [PubMed] [Google Scholar]
  • 55.Neal B, Asthagiri D, Lenhoff A. Molecular origins of osmotic second virial coefficients of proteins. Biophys J. 1998;75:2469–2477. doi: 10.1016/S0006-3495(98)77691-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Kim YC, Hummer G. Coarse-grained models for simulation of multiprotein complexes: Application to ubiquitin binding. J Mol Biol. 2008;375:1416–1433. doi: 10.1016/j.jmb.2007.11.063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Wang LP, Martinez TJ, Pande VS. Building force fields: An automatic, systematic, and reproducible approach. J Phys Chem Lett. 2014;5:1885–1891. doi: 10.1021/jz500737m. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Best RB, Zheng W, Mittal J. Balanced protein-water interactions improve properties of disordered proteins and non-specific protein association. J Chem Theor Comput. 2014;10:5113–5124. doi: 10.1021/ct500569b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Huang J, et al. CHARMM36m: An improved force field for folded and intrinsically disordered proteins. Nat Methods. 2017;14:71–73. doi: 10.1038/nmeth.4067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Robustelli P, Piana S, Shaw DE. Developing a molecular dynamics force field for both folded and disordered protein states. Proc Natl Acad Sci USA. 2018;115:E4758–E4766. doi: 10.1073/pnas.1800690115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Michalet X, Weiss S, Jäger M. Single-molecule fluorescence studies of protein folding and conformational dynamics. Chem Rev. 2006;106:1785–1813. doi: 10.1021/cr0404343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Schuler B, Hofmann H. Single-molecule spectroscopy of protein folding dynamics–expanding scope and timescales. Curr Opin Struct Biol. 2013;23:36–47. doi: 10.1016/j.sbi.2012.10.008. [DOI] [PubMed] [Google Scholar]
  • 63.Bernadó P, Svergun DI. Structural analysis of intrinsically disordered proteins by small-angle x-ray scattering. Mol Biosyst. 2012;8:151–167. doi: 10.1039/c1mb05275f. [DOI] [PubMed] [Google Scholar]
  • 64.Riback JA, et al. Innovative scattering analysis shows that hydrophobic proteins are expanded in water. Science. 2017;358:238–241. doi: 10.1126/science.aan5774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Zheng W, Best RB. An extended Guinier analysis for intrinsically disordered proteins. J Mol Biol. 2018;430:2540–2553. doi: 10.1016/j.jmb.2018.03.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Falahati H, Wieschaus E. Independent active and thermodynamic processes govern the nucleolus assembly in vivo. Proc Natl Acad Sci USA. 2017;114:1335–1340. doi: 10.1073/pnas.1615395114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Wang A, et al. A single N-terminal phosphomimic disrupts TDP-43 polymerization, phase separation, and RNA splicing. EMBO J. 2018;37:e97452. doi: 10.15252/embj.201797452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Zerze GH, Best RB, Mittal J. Sequence-and temperature-dependent properties of unfolded and disordered proteins from atomistic simulations. J Phys Chem B. 2015;119:14622–14630. doi: 10.1021/acs.jpcb.5b08619. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Debye P, Hückel E. De la theorie des electrolytes. I. Abaissement du point de congelation et phenomenes associes. Phys Z. 1923;24:185–206. [Google Scholar]
  • 70.Kapcha LH, Rossky PJ. A simple atomic-level hydrophobicity scale reveals protein interfacial structure. J Mol Biol. 2014;426:484–498. doi: 10.1016/j.jmb.2013.09.039. [DOI] [PubMed] [Google Scholar]
  • 71.Anderson JA, Lorenz CD, Travesset A. General purpose molecular dynamics simulations fully implemented on graphics processing units. J Comput Phys. 2008;227:5342–5359. [Google Scholar]
  • 72.Sugita Y, Okamoto Y. Replica-exchange molecular dynamics methods for protein folding. Chem Phys Lett. 1999;314:141–151. [Google Scholar]
  • 73.Plimpton S. Fast parallel algorithms for short-range molecular dynamics. J Comput Phys. 1995;117:1–19. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary File
Supplementary File
Supplementary File
pnas.1804177115.sm02.gif (30.7MB, gif)
Supplementary File
pnas.1804177115.sm03.gif (12.8MB, gif)
Supplementary File
Supplementary File

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES