Abstract
Intracellular membraneless organelles and their myriad cellular functions have garnered tremendous recent interest. It is becoming well accepted that they form via liquid-liquid phase separation (LLPS) of protein mixtures (often including RNA), where the organelles correspond to a protein-rich, droplet phase coexisting with a protein-poor, bulk phase. The major protein components contain disordered regions and often also RNA-binding domains, and the disordered fragments on their own easily undergo LLPS. In contrast, LLPS for structured proteins has been observed infrequently. The contrasting phase behaviors can be explained by modeling disordered and structured proteins, respectively, as polymers and colloids. These physical models also provide better understanding on the regulation of droplet formation by cellular signals and dysregulation leading to diseases.
Keywords: Membarneless organelle, phase separation, transient bonding network
MLOs: disordered proteins as drivers and myriad cellular functions
Whereas the roles of most membrane-bound organelles have been established over a century, characterization of the molecular and physical features of membraneless organelles (MLOs; see Glossary) has started only in recent years. Already, they have been associated with a wide range of biological functions, including RNA processing, ribosome biogenesis, and sequestration of mRNA (for later translation), proteins (for signaling), and compacted chromatin (for gene silencing) [1-3]. It is now well accepted that MLOs form via liquid-liquid phase separation (LLPS; Box 1), whereby they emerge from the cytoplasm or nucleoplasm as protein-rich droplets [4-8]. Like their membrane-bound counterparts, each type of MLO contains many different species of macromolecules, but the MLO identity can often be associated with one or more “driver” proteins that are most responsible for the assembly and function of the organelle [9]. For example, fibrillarin (FIB1) and nucleophosmin (NPM1) are putatively driver proteins for the subcompartmentalized organization of the nucleolus, an organelle with a primary function in ribosome pre-assembly [10]. In a number of better-characterized cases, the driver proteins always have extended, intrinsically disordered regions (IDRs), which sometimes also have RNA-binding ability (Figure 1a); quite often though, one or more separate RNA-binding domains, such as RNA-recognition motifs (RRMs), are present as well [11] (Table 1). Many studies with deletion constructs have shown that the IDRs are sufficient and perhaps necessary for LLPS [10, 12-23]. It has been speculated that intrinsic disorder may promote LLPS by enabling multiple, weakly attractive interactions [12, 14, 17, 24, 25], but the mechanism largely remains unclear.
Box 1. Phases and determination of coexistence curves.
Table 1. Disordered regions and RNA-binding domains of proteins that drive the formation of membraneless organelles.
Driver proteina | IDRsb | Enriched AAs | RNA binding | MLOs | References |
---|---|---|---|---|---|
DDX3_CAEEL (708) D0PV95 LAF-1/DDX3 |
1-202 620-708 |
9-168: R/G 624-691: G |
N-term IDR (RGG repeats) | P granule | [13, 29] |
DDX4_HUMAN (724) Q9NQI0 DDX4 |
1-257 | 58-234: G | N-term IDR? | Nuage | [17] |
FBRL_CAEEL (352) Q22053 FIB1/fibrillarin |
1-122 | 8-114: DMA/G | N-term IDR? Fibrillarin domain (119-346)? |
Nucleolus | [10] |
FUS_HUMAN (526) P35637 FUS |
1-286 356-526 |
1-165: Q/G/S/Y 166-267: G 371-526: R/G |
RRM (287-365) | Stress granule Paraspeckle DNA damage site | [12, 14, 16, 47] |
IF4F2_YEAST (914) P39936 eIF4G2 |
1-104 111-389 454-581 812-914 |
32-97: N 459-510: R/S 840-863: R/S |
MIF4G (567-810) | Stress granule | [14] |
LSM4_YEAST (187) P40070 LSm4 |
88-187 | LSM (5-81) | P body | [14] | |
NPM_XENLA (299) P07222 NPM1/nucleophosmin |
116-258 | 1-7: M 122-137: D/E 160-187: D/E |
NPM1-C (246-294) | Nucleolus | [10, 24] |
PTBP1_HUMAN (531) P26599 PTB/HNRNP I |
316-323: A only | RRM (59-143) RRM (184-260) RRM (337-411) RRM (454-529) |
Nuclear speckle | [5] | |
PUB1_YEAST (453) P32588 Pub1 |
1-64 211-330 410-453 |
8-16: Q 243-288: N 442-453: Q |
RRM (77-146) RRM (164-234) RRM (343-407) |
Stress granule | [14] |
ROA1_HUMAN (372) P09651 HNRNPA1 |
177-372 | 195-372: G | RRM (16-85) RRM (107-176) C-term IDR |
Stress granule Paraspeckle | [14, 15] |
TADBP_HUMAN (414) Q13148 TDP-43 |
258-414 | 274-413: G | RRM (106-172) RRM (193-243) |
Stress granule Paraspeckle | [15, 22] |
TIA1_HUMAN (386) P31483 TIA-1 |
331-386 | RRM (9-77) RRM (108-178) RRM (216-280) |
Stress granule | [14, 60] | |
WHI3_YEAST (661) P34761 WHI3 |
1-102 196-314 324-580 608-661 |
247-277: Q | RRM (540-613) | Stress granule | [19] |
ROA2_HUMAN (341) P22626-2 hnRNPA2 |
171-341 | 190-341: G | RRM (11-80) RRM (102-171) |
Stress granule | [18, 23] |
NPHN_HUMAN (1241) O60500 Nephrin |
1095-1171 | Nuclear body | [20] | ||
Q9TXM1_CAEEL (862) Q9TXM1 MEG-3 |
1-382 395-606 |
S | Entire length | P granule | [21] |
G5EBV6_CAEEL (693) G5EBV6 PGL-3 |
518-693 | RGG repeats (634-693) | P granule | [32] |
Each protein is identified by the UniProt entry name (with length given in parentheses) and accession number, as well as the common name.
Disorder was predicted by PONDR-VSL2 [77], with the threshold score set at 0.8 for reliability.
A hallmark of LLPS (in contrast to, e.g., protein aggregation) is thermodynamic reversibility, consistent with the liquid-like character of the droplet phase. That is, droplets can easily dissolve upon raising temperature or salt concentration and can reform when conditions are reverted. The reversibility comes because the two phases, droplet and dispersed, are in thermodynamic equilibrium (Box 1). Phase separation occurs when molecules can achieve the same low free energy by adopting two distinct types of configurations, with disparate concentrations and extents of intermolecular contacts [26]. The low free energy is achieved by distinct means in the two phases: e.g., high entropy (due to low concentration) in the dispersed phase but strong favorable (i.e., negative) enthalpy (from intermolecular contacts) in the droplet phase. Raising temperature or salt concentration leads to reduced stabilization of the droplet phase by intermolecular contacts. Beyond a critical point, the distinction between the two phases vanishes, and hence droplets dissolve. In general, the more strongly attractive the intermolecular interactions are, the wider the range of conditions for phase separation is (corresponding to raised critical temperature or critical salt concentration).
Synthetic polymers have long been known to undergo LLPS [27]. As with other thermodynamic properties of intrinsically disordered proteins (IDPs) [28], polymer models have been invoked to analyze phase boundaries. Specifically, the Flory-Huggins theory and its extensions (Box 2) have been used to fit experimental data or construct phase diagrams, providing explanations for the effects of salts, RNA, and sequence charge patterns on phase boundaries [17, 29-31]. Salt ions can screen charge-charge attraction between protein residues, and hence droplets dissolve above a critical salt concentration. Several studies have shown that the addition of RNA promoted LLPS, indicated by either reduced threshold protein concentrations for droplet formation or increased critical salt concentrations [5, 10, 14, 15, 21, 24, 32]. On the other hand, Zhang et al. [19] and Banerjee et al. [33] observed that RNA promoted LLPS only up to a point in RNA concentration, with further increase in RNA leading to LLPS suppression. In yet another variation, Wei et al. [29] recently found that, for LAF-1, a disordered driver protein for P granules [13], RNA had little effects on the threshold protein concentration and critical salt concentration, but significantly reduced the protein concentration in the droplet phase (Figure 1b). These conflicting reports were resolved in a recent theoretical study [34] (see below).
Box 2. Free energy calculation: lattice models and mean-field treatment.
Wei et al. [29] noted that the LAF-1 concentrations in the dissolved and droplet phases, approximately 0.1 and 5 mg/ml, respectively, are orders of magnitude lower than the counterparts of structured proteins including lysozyme [35] and γ-crystallins (Figure 1c,d) [36]. This contrast buttresses the notion that LLPS occurs much more easily for disordered proteins than for structured proteins, which is a focus of the present review.
LLPS of structured proteins: metastability relative to fluid-solid transition
LLPS was actually first observed on structured proteins, as a metastable step on the way to crystallization [37, 38]. The metastability (Figure 1d) can be attributed to the small ratio between the range of attractive interactions and diameter of proteins, as demonstrated on spherical models of colloidal particles (Box 3). Kinetically, the metastable droplets rich in proteins can facilitate nucleation and thereby accelerate crystallization [39-42].
Box 3. Free energy calculations: perturbation theory for colloids.
Under conditions for slow crystallization, the LLPS coexistence curves of several structured proteins have been determined, including arachin [37], lysozyme [35], γ-crystallins (Figure 1d) [36, 43], with critical protein concentrations ranged from 180 to 276 mg/ml. The critical concentrations of immunoglobulin G (IgG) antibodies were somewhat lower, at approximately 90 mg/ml, which was attributed to the nonspherical shape of the antibody molecules [44]. Recently a peptide oligomer was found to phase separate with an even lower critical concentration of 50 mg/ml; loose packing within the oligomer was suggested as an explanation [45].
The high concentrations required pose one difficulty for observing LLPS of structured proteins. Structured proteins also tend to have low critical temperatures (Tc). When Tc is below the freezing point of the protein solution, it is not possible to directly observe LLPS. Crowding agents like polyethylene glycol (PEG) have been found to promote LLPS, raising Tc to above the freezing point [43, 45]. By extrapolating, one can deduce the coexistence curve for the protein in the absence of crowding.
Transient bonding networks: common organizational principle for protein droplets
So far there has been very little crosstalk between the growing community of scientists working on LLPS of MLO-driving disordered proteins and the more established counterpart on LLPS of structured proteins. Connecting the research in the two communities will help us establish a common physical basis for phase separation of proteins, exchange knowledge between the two communities, and identify what are unique about IDPs and about structured proteins.
Droplet formation requires a sufficient extent of attraction between protein molecules to provide stability. On the other hand, to be in a liquid phase, individual interactions between two molecules must be easily breakable and hence relatively weak. Therefore, by necessity, proteins, whether structured or disordered, in the droplet phase form weakly attractive interactions with multiple partners, resulting in transient bonding networks (Figure 2a) [46].
For rigid structured proteins, it has become possible to use an all-atom representation in simulating phase separation and calculating coexistence curves (Figure 2) [46]. These simulations provide an atomistic view of the bonding networks in protein droplets. With atomistic calculations of coexistence curves, it is possible to test whether ideas such as nonspherical shape and loose packing are correct in explaining low critical concentrations for some structured proteins. Moreover, even highly homologous γ-crystallins can differ in Tc in excess of 30 °C (Figure 1d). The calculations afford an opportunity to relate changes in phase behavior to changes in amino acid sequence. Such relations may provide valuable insight for understanding how mutations (e.g., ones associated with diseases) and posttranslational modifications affect LLPS of MLO-driving proteins (see below).
In addition to changes in amino acid sequence, the presence of other macromolecules can also influence the bonding networks of droplet-forming proteins and hence their phase behaviors. As noted, PEG has been used to raise Tc for structured proteins [43, 45]. Similarly, crowding agents such as PEG, Ficoll, and dextran have been found to promote LLPS of MLO-driving proteins [14, 15, 47].
Crowders promote LLPS by preferentially partitioning in the dispersed phase, and therefore displacing proteins into the droplet phase to strengthen bonding networks there [34]. A macromolecular component, e.g., another protein or RNA, starts to partition in the droplet phase if it has attraction for the droplet-forming protein. When the macromolecule-protein attraction is mild, the presence of the macromolecule in the droplet phase serves to disrupt the bonding networks of the protein and thereby suppress LLPS. Such an effect was observed for human and bovine serum albumin, respectively, on the phase separation of an IgG antibody [44] and hnRNPA1 [48]. On the other hand, when the macromolecule-protein attraction is stronger than the protein-protein attraction, the macromolecule can strengthen the protein bonding networks and thus promote LLPS. This is likely the reason for the effects of RNA in several studies [5, 10, 14, 15, 21, 24, 32]. However, the promotional effects persist only up to a certain RNA-to-protein molar ratio, beyond which RNA displaces too much of the protein from the droplet phase and thereby again disrupts the bonding networks. Such dual effects of RNA have been observed by Zhang et al. [19] and Banerjee et al. [33].
It should be noted that MLO-driving proteins often contain both IDRs and structured domains (including those binding RNA; see Table 1). Several studies have shown that IDRs and structured domains can act synergistically in promoting phase separation [24, 48-50].
Physical basis for contrasting phase behaviors of disordered and structured proteins
A common physical basis notwithstanding, structured and disordered proteins nevertheless exhibit very different phase behaviors, in terms of high Tc and low critical concentrations for the latter proteins and the general ease in observing LLPS for them. Whereas structured proteins can form crystals and LLPS for them is often metastable relative to the fluid-solid transition, disordered proteins cannot form crystals and hence do not have a similar fluid-solid transition for competition [26]. Disordered proteins can form solid-like condensates such as gels and fibrils.
The contrast in Tc and critical concentration between structured and disordered proteins is reminiscent of the differing phase behaviors of colloid particles (Box 3) and polymer chains (Box 4). We represent colloid or polymer concentration by the volume fraction ϕ. Relative to a colloid that models globular proteins, a 100-residue polymer, which could serve as a model for IDPs, has a Tc that is 4.2 times higher and a critical concentration (ϕc) that is 5.1 times lower. How can we understand these differences?
Box 4. Free energy calculations: perturbation theory for polymers.
At phase equilibrium, the dispersed and droplet phases must have equal chemical potentials. The chemical potential can be decomposed,
[1] |
The first term on the right-hand side is the ideal part, which is the chemical potential if intermolecular interactions were totally absent, given by
[2] |
The second term, μrep, is the contribution from steric repulsion between colloids or polymers; this would be the only contribution from intermolecular interactions if the molecules were purely repulsive toward each other. This term is positive and an increasing function of ϕ, reflecting the difficulty in inserting a molecule into an already crowded solution. The third term, μatt, is the contribution from intermolecular interactions beyond steric repulsion, and does not have to be purely enthalpic. The μid and μrep terms favor the dispersed phase, whereas μatt favors the droplet phase and must become sufficiently negative at increasing ϕ for LLPS to occur. (Note that, when crowders are present, they preferentially partition into the dispersed phase and raise μrep there; then μatt does not need to be as negative as in the absence of crowders.) In general Tc increases with increasing magnitude of μatt.
In Figure 3, Key Figure we compare the component and total chemical potentials of the colloid and 100-residue polymer. Each monomer unit of the polymer can have attractive interactions with all other monomers (in the same chain or in other chains), in a way similar to attractive interactions between colloid particles. Therefore μatt for the polymer accumulates over the monomer units, leading to a much higher Tc than that of the colloid. Due to the same reason, the chemical potential results for the polymer are shown in Figure 3b on a per monomer basis (i.e., divided by the chain length L), and at a temperature that is 4 times higher than for the colloid. For the colloid, μid has a rapid rise near ϕ = 0, and subsequently, μrep takes off; μatt catches up with μid + μrep only at a high ϕ. That explains why the concentrations of colloids (and, by inference, globular proteins) in the droplet phase are high. The concentrations of colloids in the dispersed phase cannot be too low either, since that would make the dispersed phase too stable (due to a strongly favorable μid) compared with the droplet phase.
In contrast, for the polymer, on a per monomer basis, μid/L is small, and the rise of μrep/L is much less rapid than μrep for a colloid particle (the two would be the same were the monomer units not connected into a chain). Therefore μatt/L can catch up with μid/L + μrep/L at relatively low ϕ. That explains why the concentrations of polymers (and, by inference, IDPs) in the droplet phase are low.
The polymer model also predicts a much higher Tc than the colloid model (see Box 3 Figure I and Box 4 Figure I). This difference is related to the fact that each monomer unit in a polymer chain has the freedom to interact with all the monomer units in any other chain. Therefore the flexible polymer chains, as well as IDPs, can easily form multivalent interactions.
Regulation of MLO formation by cellular signals
The foregoing presentation makes it easy to explain how various cellular signals can regulate protein droplet formation. Environmental stress can be transmitted into changes in intracellular osmotic pressure or temperature [17, 51], and these basic physical changes can bring proteins into or out of LLPS conditions. Effects of varying salt concentrations and pH on LLPS, by modulating the magnitude of μatt, have been amply demonstrated in in vitro studies, and are likely exploited in vivo. Interestingly, cells may also perturb μrep, by changing the level of macromolecular crowding through an increase or decrease in cell size [52].
LLPS occurs below the critical point but only when protein concentrations are above a threshold. Therefore intracellular protein concentration provides another means for regulating MLO formation [15]. RNA in particular can promote phase separation, but only within a certain range, and hence RNA expression level can also control whether MLOs form. Molecules that compete for RNA binding with droplet-forming proteins can limit the amount of RNA that enter into droplets and thus provide yet another layer of regulation [21, 32].
Posttranslational modifications (PTMs) regulate molecular structural organizations at all levels, and MLOs are no exception. The IDRs of droplet-forming proteins provide ready accessibility for PTMs, and the PTMs can promote or suppress LLPS by enabling or disrupting intermolecular interactions. For example, Tyr phosphorylations of the nephrin C-terminal IDR enable multivalent binding with Nck SH2 domains, and thereby promote LLPS of the Nck/N-WASP system [5]. On the other hand, Ser/Thr phosphorylation and phosphomimetic (Ser/Thr to Glu) mutation of the FUS N-terminal IDR disrupt transient intermolecular interactions and suppress LLPS [53]. Similarly, Ser phosphorylation of the MEG group of IDPs promotes P granule disassembly whereas dephosphorylation promotes granule assembly [54]. Arg methylations of both the hnRNPA2 C-terminal IDR [23] and the DDX4 N-terminal IDR [17] were found to suppress LLPS. An alternative splicing of the DDX4 N-terminal IDR also abrogated LLPS [17].
Droplets of structured proteins can nucleate crystal growth and have been associated with the genesis of protein condensation diseases such as cataract [38, 55, 56] and sickle cell anemia [57, 58]. Similarly, LLPS of disordered proteins can facilitate further transition to solid-like condensates such as gels and fibrils, either as normal cellular response [59] or for pathogenesis [15, 22, 23, 47, 53, 60]. Disease-associated mutations may either promote LLPS or accelerate liquid-to-solid transition.
Concluding remarks
The physical basis for the LLPS of proteins, both structured and disordered, is becoming clear. Proteins are driven into the dispersed and droplet phases by different forces: translational freedom of the individual molecules and steric repulsion between them for the dispersed phase whereas intermolecular attraction for the droplet phase. It is important to recognize that the phase equilibrium is determined by the balance of steric (or crowding) and attractive interactions. Macromolecular crowders, by preferentially partitioning into the dispersed phase, can have significant effects on LLPS, as demonstrated in computational and in vitro experimental studies, and very likely on in vivo MLO formation. Other macromolecules, in particular RNA, can partition into the droplet phase, thereby strengthening or disrupting transient bonding networks there and affecting the phase equilibrium.
Macromolecular crowders and RNA are respective examples of species that are excluded from and recruited into protein droplets. MLOs contain multiple components and are surrounded by many more non-component species. Moreover, different types of MLOs may be immiscible and may even organize into core-shell structures (droplet inside droplet) [10]. A fruitful direction would be to characterize the differential partitioning of various species and their effects on MLO assembly and disassembly (see Outstanding Questions).
Outstanding Questions.
Disordered proteins drive the formation of membraneless organelles (MLOs), but many other components help regulate the assembly/disassembly and functions of MLOs. The recruitment of certain macromolecular species into MLOs and the exclusion of many others, and the consequence on MLO assembly/disassembly are only starting to be understood in terms of intermolecular interactions. Polymer crowding agents and RNA serve as extreme examples, but the effects of the differential partitioning of many more species, their mutual influences, and their individual and collective effects on MLO structure and function are yet to be well characterized.
Transient bonding networks appear to be a general mechanism for the internal structural organization of MLOs. Expanding the knowledge in this area is potentially a new frontier for structural biology and computational biophysics.
Liquid-liquid coexistence curves have been determined experimentally for some proteins and more should be done. Even more urgently, calculations of coexistence curves, based on realistic modeling of proteins and rigorous implementation of thermodynamic principles, can bring tremendous physical insights to, in particular, the effects of posttranslational modifications and disease-associated mutations, but will require solutions to technical challenges.
Pathogenesis of MLO proteins often involves transition to solid-like condensates. Physical models are needed to gain a foundational understanding of this process.
Our knowledge about transient bonding networks inside protein droplets is still scant. Here computation can well complement structural biology techniques. For structured proteins, computation at the all-atom level is already feasible [46], but for disordered proteins, conformational flexibility has limited computation to a coarsegrained level [23]. For subtle effects of PTMs and disease-causing mutations, an all-atom representation may be necessary for quantitative prediction. All-atom computation for IDP LLPS will be a great challenge for the future.
Transitions of IDP droplets into solid-like condensates (e.g., gels and fibrils) are implicated both in normal cellular responses and in diseases. A general physical model for such transitions is yet to be developed, in part because they often appear irreversible. Once again one may have to go to the colloid and polymer literature for some guidance.
Highlights.
Phase separation leading to protein droplets and membraneless organelles (MLOs) is determined by the balance among translational entropy of protein molecules and steric and attractive interactions between molecules.
Colloids and polymers serve as good models for understanding the different phase behaviors of structured and disordered proteins. Disordered proteins are characterized by both extensive attraction and low energetic cost from steric repulsion, contributing to easy observation of phase separation.
Protein molecules form transient bonding networks in the droplet phase. The networks can be strengthened or disrupted by posttranslational modifications and disease-associated mutations and by other macromolecular species such as RNA, leading to assembly and disassembly of MLOs.
Acknowledgments
This work was supported by National Institutes of Health Grant GM118091.
Glossary
- Coexistence curve
a relation between two thermodynamic properties, such as temperature and pressure or temperature and concentration, that is followed when two phases are at equilibrium. Also known as phase boundary.
- Critical point
a unique thermodynamic condition at which distinction between two fluid phases (gas and liquid for simple molecular fluids or dispersed and droplet for macromolecular solutions) vanishes. In particular, the concentrations of the two phases become identical and are known as the critical concentration. Phase separation occurs only on one side of the critical point, e.g., below the critical temperature.
- IDR
intrinsically disordered region
- IDP
intrinsically disordered protein
- LLPS
liquid-liquid phase separation
- MLO
membraneless organelle
- PTM
posttranslational modification
- RRM
RNA-recognition motif, an amino-acid sequence motif that consists of approximately 90 residues and forms a structural fold with a four-stranded β-sheet packed against two α-helices
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- 1.Mitrea DM, Kriwacki RW. Phase separation in biology; functional organization of a higher order. Cell Commun Signal. 2016;14:1. doi: 10.1186/s12964-015-0125-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Larson AG, et al. Liquid droplet formation by HP1alpha suggests a role for phase separation in heterochromatin. Nature. 2017;547:236–240. doi: 10.1038/nature22822. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Strom AR, et al. Phase separation drives heterochromatin domain formation. Nature. 2017;547:241–245. doi: 10.1038/nature22989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.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]
- 5.Li P, et al. Phase transitions in the assembly of multivalent signalling proteins. Nature. 2012;483:336–340. doi: 10.1038/nature10879. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Hyman AA, et al. Liquid-liquid phase separation in biology. Annu Rev Cell Dev Biol. 2014;30:39–58. doi: 10.1146/annurev-cellbio-100913-013325. [DOI] [PubMed] [Google Scholar]
- 7.Brangwynne CP, et al. Polymer physics of intracellular phase transitions. Nat Phys. 2015;11:899–904. [Google Scholar]
- 8.Qin S, Zhou HX. Protein folding, binding, and droplet formation in cell-like conditions. Curr Opin Struct Biol. 2017;43:28–37. doi: 10.1016/j.sbi.2016.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Aguilera-Gomez A, Rabouille C. Membrane-bound organelles versus membrane-less compartments and their control of anabolic pathways in Drosophila. Dev Biol. 2017;428:310–317. doi: 10.1016/j.ydbio.2017.03.029. [DOI] [PubMed] [Google Scholar]
- 10.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]
- 11.Courchaine EM, et al. Droplet organelles? EMBO J. 2016;35:1603–1612. doi: 10.15252/embj.201593517. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Burke KA, et al. 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.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 U S A. 2015;112:7189–7194. doi: 10.1073/pnas.1504822112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Lin Y, et al. Formation and maturation of phase-separated liquid droplets by RNA-binding proteins. Mol Cell. 2015;60:208–219. doi: 10.1016/j.molcel.2015.08.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.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]
- 16.Murakami T, et al. ALS/FTD mutation-induced phase transition of FUS liquid proplets and reversible hydrogels into irreversible hydrogels impairs RNP granule function. Neuron. 2015;88:678–690. doi: 10.1016/j.neuron.2015.10.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Nott TJ, et al. Phase transition of a disordered nuage protein generates environmentally responsive membraneless organelles. Mol Cell. 2015;57:936–947. doi: 10.1016/j.molcel.2015.01.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.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]
- 19.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]
- 20.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]
- 21.Smith J, et al. Spatial patterning of P granules by RNA-induced phase separation of the intrinsically-disordered protein MEG-3. eLife. 2016;5:e21337. doi: 10.7554/eLife.21337. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Conicella AE, et al. 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]
- 23.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 doi: 10.1016/j.molcel.2017.12.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Mitrea DM, et al. Nucleophosmin integrates within the nucleolus via multi-modal interactions with proteins displaying R-rich linear motifs and rRNA. eLife. 2016;5:e13571. doi: 10.7554/eLife.13571. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Uversky VN. Intrinsically disordered proteins in overcrowded milieu: Membrane-less organelles, phase separation, and intrinsic disorder. Curr Opin Struct Biol. 2016;44:18–30. doi: 10.1016/j.sbi.2016.10.015. [DOI] [PubMed] [Google Scholar]
- 26.Zhou HX, Pang X. Electrostatic interactions in protein structure, folding, binding, and condensation. Chem Rev. 2018;118:1691–1741. doi: 10.1021/acs.chemrev.7b00305. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Flory PJ. Thermodynamics of high polymer solutions. J Chem Phys. 1942;10:51–61. [Google Scholar]
- 28.Zhou HX. Polymer models of protein stability, folding, and interactions. Biochemistry. 2004;43:2141–2154. doi: 10.1021/bi036269n. [DOI] [PubMed] [Google Scholar]
- 29.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]
- 30.Chang LW, et al. Sequence and entropy-based control of complex coacervates. Nat Commun. 2017;8:1273. doi: 10.1038/s41467-017-01249-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.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]
- 32.Saha S, et al. Polar positioning of phase-separated liquid compartments in cells regulated by an mRNA competition mechanism. Cell. 2016;166:1572–1584 e1516. doi: 10.1016/j.cell.2016.08.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Banerjee PR, et al. Reentrant phase transition drives dynamic substructure formation in ribonucleoprotein droplets. Angew Chem Int Ed Engl. 2017;56:11354–11359. doi: 10.1002/anie.201703191. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Nguemaha V, Zhou HX. Liquid-liquid phase separation of patchy particles illustrates disparate effects of constituents on protein droplets. Sci Rep. 2018 doi: 10.1038/s41598-018-25132-1. submitted. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Taratuta VG, et al. Liquid-liquid phase separation of aqueous lysozyme solutions: effects of pH and salt identity. J Phys Chem. 1990;94:2140–2144. [Google Scholar]
- 36.Broide ML, et al. Binary-liquid phase separation of lens protein solutions. Proc Natl Acad Sci U S A. 1991;88:5660–5664. doi: 10.1073/pnas.88.13.5660. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Tombs MP, et al. Protein solubility: Phase separation in arachin-salt-water systems. Int J Peptide Protein Res. 1974;6:253–277. doi: 10.1111/j.1399-3011.1974.tb02384.x. [DOI] [PubMed] [Google Scholar]
- 38.Tanaka T, et al. Phase separation of a protein-water mixture in cold cataract in the young rat lens. Science. 1977;197:1010–1012. doi: 10.1126/science.887936. [DOI] [PubMed] [Google Scholar]
- 39.ten Wolde PR, Frenkel D. Enhancement of protein crystal nucleation by critical density fluctuations. Science. 1997;277:1975–1978. doi: 10.1126/science.277.5334.1975. [DOI] [PubMed] [Google Scholar]
- 40.Gliko O, et al. A metastable prerequisite for the growth of lumazine synthase crystals. J Am Chem Soc. 2005;127:3433–3438. doi: 10.1021/ja043218k. [DOI] [PubMed] [Google Scholar]
- 41.Veesler S, et al. Crystallization in the presence of a liquid- liquid phase separation. Org Process Res Dev. 2006;10:841–845. [Google Scholar]
- 42.Schubert R, et al. Real-time observation of protein dense liquid cluster evolution during nucleation in protein crystallization. Cryst Growth Des. 2017;17:954–958. [Google Scholar]
- 43.Annunziata O, et al. Observation of liquid-liquid phase separation for eye lens gammaS-crystallin. Proc Natl Acad Sci U S A. 2003;100:970–974. doi: 10.1073/pnas.242746499. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Wang Y, et al. Phase separation in solutions of monoclonal antibodies and the effect of human serum albumin. Proc Natl Acad Sci U S A. 2011;108:16606–16611. doi: 10.1073/pnas.1112241108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Wang Y, et al. Liquid-liquid phase separation in oligomeric peptide solutions. Langmuir. 2017;33:7715–7721. doi: 10.1021/acs.langmuir.7b01693. [DOI] [PubMed] [Google Scholar]
- 46.Qin S, Zhou HX. Fast method for computing chemical potentials and liquid-liquid phase equilibria of macromolecular solutions. J Phys Chem B. 2016;120:8164–8174. doi: 10.1021/acs.jpcb.6b01607. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.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]
- 48.Protter DSW, et al. Intrinsically disordered regions can contribute promiscuous interactions to RNP granule assembly. Cell Rep. 2018;22:1401–1412. doi: 10.1016/j.celrep.2018.01.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Banjade S, et al. Conserved interdomain linker promotes phase separation of the multivalent adaptor protein Nck. Proc Natl Acad Sci U S A. 2015;112:E6426–6435. doi: 10.1073/pnas.1508778112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Harmon TS, et al. Intrinsically disordered linkers determine the interplay between phase separation and gelation in multivalent proteins. Elife. 2017;6 doi: 10.7554/eLife.30294. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Rabouille C, Alberti S. Cell adaptation upon stress: the emerging role of membrane-less compartments. Curr Opin Cell Biol. 2017;47:34–42. doi: 10.1016/j.ceb.2017.02.006. [DOI] [PubMed] [Google Scholar]
- 52.Joyner RP, et al. A glucose-starvation response regulates the diffusion of macromolecules. eLife. 2016;5:e09376. doi: 10.7554/eLife.09376. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Monahan Z, et al. Phosphorylation of the 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]
- 54.Wang JT, et al. Regulation of RNA granule dynamics by phosphorylation of serine-rich, intrinsically disordered proteins in C. elegans. eLife. 2014;3:e04591. doi: 10.7554/eLife.04591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Kmoch S, et al. Link between a novel human gamma D-crystallin allele and a unique cataract phenotype explained by protein crystallography. Hum Mol Genet. 2000;9:1779–1786. doi: 10.1093/hmg/9.12.1779. [DOI] [PubMed] [Google Scholar]
- 56.Pande A, et al. Crystal cataracts: human genetic cataract caused by protein crystallization. Proc Natl Acad Sci U S A. 2001;98:6116–6120. doi: 10.1073/pnas.101124798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.San Biagio PL, Palma MU. Spinodal lines and Flory-Huggins free-energies for solutions of human hemoglobins HbS and HbA. Biophys J. 1991;60:508–512. doi: 10.1016/S0006-3495(91)82078-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Galkin O, et al. Liquid-liquid separation in solutions of normal and sickle cell hemoglobin. Proc Natl Acad Sci U S A. 2002;99:8479–8483. doi: 10.1073/pnas.122055299. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Munder MC, et al. A pH-driven transition of the cytoplasm from a fluid- to a solid-like state promotes entry into dormancy. eLife. 2016;5:e09347. doi: 10.7554/eLife.09347. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Mackenzie IR, et al. TIA1 mutations in amyotrophic lateral sclerosis and frontotemporal dementia promote phase separation and alter stress granule dynamics. Neuron. 2017;95:808–816. doi: 10.1016/j.neuron.2017.07.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Lee TD, Yang CN. Statistical theory of equations of state and phase transitions. II. Lattice gas and Ising model. Phys Rev. 1952;87:410–419. [Google Scholar]
- 62.Muthukumar M. Thermodynamics of polymer solutions. J Chem Phys. 1986;85:4722–4728. [Google Scholar]
- 63.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]
- 64.Simon JR, et al. 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]
- 65.Ambadipudi S, et al. Liquid-liquid phase separation of the microtubule-binding repeats of the Alzheimer-related protein Tau. Nat Commun. 2017;8:275. doi: 10.1038/s41467-017-00480-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.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. e1019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Carnahan NF, Starling KE. Equation of state for nonattracting rigid spheres. J Chem Phys. 1969;51:635–636. [Google Scholar]
- 68.Barker JA, Henderso D. Perturbation theory and equation of state for fluids: The square-well potential. J Chem Phys. 1967;47:2856–2861. [Google Scholar]
- 69.Chang J, Sandler SI. A real function representation for the structure of the hard-sphere fluid. Mol Phys. 1994;81:735–744. [Google Scholar]
- 70.Vliegenthart GA, Lekkerkerker HNW. Predicting the gas-liquid critical point from the second virial coefficient. J Chem Phys. 2000;112:5364–5369. [Google Scholar]
- 71.Velasco E, et al. Phase diagram of colloidal systems. Langmuir. 1998;14:5652–5655. [Google Scholar]
- 72.Gast AP, et al. Polymer-induced phase separations in non-aqueous colloidal suspensions. J Colloid Interf Sci. 1983;96:251–267. [Google Scholar]
- 73.Choi Y, et al. Hard-sphere radial distribution functions for face-centered cubic and hexagonal close-packed phases: Representation and use in a solid-state perturbation theory. J Chem Phys. 1991;95:7548–7561. [Google Scholar]
- 74.Wertheim MS. Thermodynamic perturbation-theory of polymerization. J Chem Phys. 1987;87:7323–7331. [Google Scholar]
- 75.Escobedo FA, De Pablo JJ. Chemical-potential and equations of state of hard-core chain molecules. J Chem Phys. 1995;103:1946–1956. [Google Scholar]
- 76.Escobedo FA, De Pablo JJ. Simulation and prediction of vapour-liquid equilibria for chain molecules. Mol Phys. 1996;87:347–366. [Google Scholar]
- 77.Peng K, et al. Length-dependent prediction of protein intrinsic disorder. BMC Bioinformatics. 2006;7:208. doi: 10.1186/1471-2105-7-208. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Berland CR, et al. Solid-liquid phase boundaries of lens protein solutions. Proc Natl Acad Sci U S A. 1992;89:1214–1218. doi: 10.1073/pnas.89.4.1214. [DOI] [PMC free article] [PubMed] [Google Scholar]