Fact-checked by Grok 2 weeks ago

Phase separation

Phase separation is a fundamental in which a homogeneous or spontaneously segregates into two or more distinct phases with differing compositions, densities, and physical properties, driven by the system's tendency to minimize its . This phenomenon occurs when the mixture becomes thermodynamically unstable, often due to changes in , , concentration, or external fields, leading to phase transitions that can be either (involving and growth) or continuous (via ). Phase separation is ubiquitous in nature and technology, manifesting in everyday examples like the separation of oil and or the demixing of alloys upon cooling. In physics and , phase separation is central to understanding the behavior of multicomponent systems, such as fluids, blends, and metallic alloys, where it influences microstructure formation and material properties like strength and . Mechanisms like , characterized by an amplified instability in composition fluctuations, enable rapid phase ordering without energy barriers, contrasting with nucleation processes that require overcoming an activation energy for droplet formation. These dynamics are described by classical theories, including Cahn-Hilliard equations, which model conserved order parameters in diffusive systems. In chemistry, phase separation underpins applications in polymer processing, such as thermally induced phase separation () for creating porous scaffolds in , and in colloidal stability where prevent unwanted demixing in emulsions. More recently, in , liquid-liquid phase separation (LLPS) has gained prominence as a mechanism for intracellular organization, enabling the formation of membraneless compartments like nucleoli and stress granules through multivalent interactions of proteins and ; as of 2025, research continues to uncover its roles in stem cell biology, tumorigenesis, and metabolic regulation. These biomolecular condensates concentrate molecules to enhance reaction rates and signaling, with implications for cellular , development, and diseases such as neurodegeneration when dysregulated.

Fundamentals

Definition and Overview

Phase separation is the by which a previously homogeneous single-phase becomes unstable and spontaneously divides into multiple coexisting phases exhibiting distinct , densities, or structures, typically induced by changes in external conditions such as , , or overall . This demixing occurs because the separated state possesses a lower compared to the uniform mixture, as the seeks to minimize its total free energy G = H - TS, where H is the , T is the absolute , and S is the . The driving force stems from the competition between enthalpic contributions from intermolecular attractions or repulsions and entropic effects from molecular configurations, leading to macroscopic phase domains when the free energy curve for mixing develops regions of negative convexity. The conceptual origins of phase separation trace back to the late , with foundational theoretical advancements in the of fluids and heterogeneous systems. In 1873, introduced his , which incorporated corrections for molecular volume and attractive forces, enabling the first quantitative prediction of phase coexistence and the liquid-gas transition in real gases. Building on this, J. Willard Gibbs developed the in his landmark 1876–1878 publication On the Equilibrium of Heterogeneous Substances, providing a general framework for determining the conditions under which multiple phases can stably coexist in a . The Gibbs phase rule quantifies the variability of multiphase equilibria and is given by F = C - P + 2, where F represents the (the number of intensive variables, such as , , or , that can be independently varied without changing the number or nature of phases in ), C is the number of independently variable components, and P is the number of phases. This relation arises from balancing the total number of variables against the constraints imposed by conditions. For a general multicomponent without chemical reactions, the intensive variables include T, p, and C-1 independent variables (e.g., mole fractions) per phase, yielding a total of P(C + 1) - P = PC + P - P = PC independent variables across P phases (accounting for the sum-to-unity constraint per phase). requires T and p to be uniform (imposing P-1 constraints each), and the of each component to be equal across all phases (imposing C(P-1) constraints total), resulting in $2(P-1) + C(P-1) = (C + 2)(P - 1) constraints. Thus, F = PC - (C + 2)(P - 1) = PC - CP + C - 2P + 2 = C - P + 2. In a (C = 2), this simplifies to F = 4 - P; for example, with two phases (P = 2), F = 2, allowing independent specification of T and one variable (e.g., at fixed pressure) to define the state uniquely. Illustrative examples of phase separation abound in everyday and materials contexts. A classic case is the spontaneous separation of oil and water into immiscible layers, where hydrophobic effects drive the nonpolar oil molecules to aggregate, minimizing unfavorable water-oil contacts and thus the system's . Similarly, in polymer blends, incompatible macromolecules like and phase separate into micron-scale domains upon cooling from a melt, as entropic mixing penalties outweigh weak intermolecular attractions, yielding materials with tailored mechanical properties. In phase diagrams, boundaries such as the and spinodal curves demarcate regions of thermodynamic from those prone to separation.

Thermodynamic Principles

Phase separation in mixtures is governed by the principles of , particularly the minimization of the G. At , the chemical potentials \mu_i of each component i must be equal across coexisting phases, ensuring that the system achieves the lowest possible state. This condition arises from the fundamental relation dG = -SdT + VdP + \sum \mu_i dn_i, where for a at constant and , the requires \mu_i^\alpha = \mu_i^\beta for phases \alpha and \beta. For binary mixtures, the common tangent construction on the -composition curve illustrates this: the tangent line connecting the free energy curves of two phases touches them at points of equal chemical potential (slopes \partial G / \partial x = \mu_1 - \mu_2), defining the compositions of the equilibrium phases. Binary temperature-composition phase diagrams map the equilibrium conditions for phase separation, with the x-axis representing composition (e.g., mole fraction x) and the y-axis temperature at constant pressure. The binodal curve separates single-phase and two-phase regions, constructed by identifying common tangents on the Gibbs free energy curves at varying temperatures; tie lines connect the equilibrium compositions of coexisting phases at a given temperature. Within the two-phase region, the lever rule determines the relative phase fractions: for an overall composition x, the fraction of phase \alpha is f_\alpha = (x_\beta - x)/(x_\beta - x_\alpha), and f_\beta = 1 - f_\alpha, reflecting the conservation of mass and minimization of total free energy along the tie line. A key model for phase separation in polymer systems is the Flory-Huggins theory, which provides the molar free energy of mixing as \frac{\Delta G_\text{mix}}{NkT} = \frac{\phi}{N} \ln \phi + (1 - \phi) \ln (1 - \phi) + \chi \phi (1 - \phi), where \phi is the volume fraction of polymer, N is the degree of polymerization, k is Boltzmann's constant, T is temperature, and \chi is the Flory interaction parameter capturing enthalpic contributions. Phase separation occurs when \chi > 2/N for large N, leading to a miscibility gap; the entropy term favors mixing, while the \chi term promotes demixing for poor solvents (\chi > 0.5). The critical point marks the temperature and composition where the two phases become indistinguishable, located at the top of the miscibility gap where the binodal and spinodal curves meet. At this point, the third and fourth derivatives of the free energy with respect to composition vanish, signaling the onset of phase separation. Universality near the critical point is described using reduced variables, such as reduced temperature t = (T_c - T)/T_c and order parameter, allowing critical exponents to be independent of microscopic details across systems. Regions of the are classified by the curvature of the : the single-phase region is stable where \partial^2 G / \partial x^2 > 0, the metastable region (between and spinodal) has \partial^2 G / \partial x^2 > 0 but allows , and the unstable region (inside the spinodal) has \partial^2 G / \partial x^2 < 0, where small fluctuations spontaneously amplify.

Mechanisms

Binodal Decomposition

Binodal decomposition refers to the process of phase separation that occurs in the metastable region of a binary mixture, where the system is located between the and the spinodal boundary in the phase diagram. In this regime, phase separation proceeds through a nucleation and growth mechanism, requiring the system to overcome a free energy barrier to form stable nuclei of the new phase, in contrast to the barrierless process within the . The binodal curve represents the locus of points in the phase diagram where two phases coexist in thermodynamic equilibrium, delineating the boundary between the single-phase region and the two-phase coexistence region. For compositions inside the binodal but outside the spinodal, the homogeneous mixture is metastable, meaning it is locally stable but not at the global minimum free energy, thus necessitating an activation process for decomposition. This curve is determined by the equality of chemical potentials and pressures between the coexisting phases, as derived from the common tangent construction on the free energy curve. The kinetics of binodal decomposition are described by classical nucleation theory (CNT), which posits that the formation of a new phase begins with the creation of small clusters or embryos whose size must exceed a critical radius to grow stably. The free energy change associated with forming a spherical nucleus of radius r is given by \Delta G(r) = \frac{4}{3} \pi r^3 \Delta \mu + 4 \pi r^2 \sigma, where \Delta \mu is the chemical potential difference driving the phase change (related to supersaturation) and \sigma is the interfacial tension between phases. The maximum free energy barrier \Delta G^* occurs at the critical radius r^* = -\frac{2\sigma}{\Delta \mu}, with \Delta G^* = \frac{16 \pi \sigma^3}{3 (\Delta \mu)^2}. This barrier height decreases with increasing supersaturation (larger |\Delta \mu|), making nucleation more probable deeper in the metastable region. The nucleation rate is then exponentially sensitive to \Delta G^*, as J \propto \exp\left( -\frac{\Delta G^*}{kT} \right), where k is Boltzmann's constant and T is temperature. Following nucleation, the subsequent growth of domains occurs through diffusion-limited attachment of material to the nuclei, leading to a coarsening process where larger domains grow at the expense of smaller ones to minimize interfacial energy. This late-stage coarsening is governed by the , which predicts that the average domain size R scales as R \sim t^{1/3} in three dimensions for diffusion-controlled , assuming a dilute dispersion of spherical precipitates. The theory, developed independently by in 1958 and in 1961, derives this scaling from the continuity equation for solute concentration and the , which relates solubility to curvature. Key factors influencing binodal decomposition include the depth of undercooling or supersaturation, which controls the nucleation barrier and rate, and the minimization of interfacial energy, which drives the coarsening dynamics. Greater undercooling reduces \Delta G^*, accelerating the process, while lower interfacial tension facilitates both nucleation and growth. A representative example of binodal decomposition is the nucleation of liquid droplets from a supersaturated vapor, such as water vapor in air, where clusters form and grow into fog or cloud droplets once exceeding the critical size determined by CNT. In such systems, experiments with argon vapor at elevated supersaturations have confirmed the liquid-like nature of critical nuclei and the subsequent diffusion-limited growth.

Spinodal Decomposition

Spinodal decomposition represents a barrierless mechanism of phase separation that occurs within the spinodal region of the phase diagram, where the homogeneous mixture is thermodynamically unstable, allowing infinitesimal composition fluctuations to grow spontaneously through diffusion without an energy barrier. This process contrasts with nucleation outside the spinodal, as it involves no critical nucleus formation and leads to the development of interconnected domains rather than isolated droplets. The spinodal curve delineates the boundary of this unstable region and is defined as the locus of points where the second derivative of the G with respect to the composition x vanishes, i.e., \partial^2 G / \partial x^2 = 0. Inside the spinodal, the negative curvature of the free energy surface (\partial^2 G / \partial x^2 < 0) renders the system susceptible to phase separation, with thermal fluctuations amplifying into macroscopic domains over time. The dynamics of spinodal decomposition are governed by the Cahn-Hilliard equation, a conserved order parameter model that describes diffusive transport driven by chemical potential gradients: \frac{\partial \phi}{\partial t} = \nabla \cdot \left[ M \nabla \frac{\delta F}{\delta \phi} \right], where \phi is the concentration field serving as the order parameter, M is the mobility (assumed constant), and F[\phi] is the Ginzburg-Landau free energy functional incorporating both bulk and gradient contributions: F[\phi] = \int \left[ f(\phi) + \frac{\kappa}{2} |\nabla \phi|^2 \right] d\mathbf{r}. Here, f(\phi) is the local homogeneous free energy density (often modeled as a double-well potential), and \kappa > 0 is the gradient energy coefficient that penalizes sharp interfaces. Linear stability analysis of the uniform state reveals the growth rate \omega(\mathbf{k}) of Fourier modes with wavenumber k = |\mathbf{k}|: \omega(k) = -M k^2 \left( \frac{\partial^2 f}{\partial \phi^2} + \kappa k^2 \right). Within the spinodal, \partial^2 f / \partial \phi^2 < 0, so long-wavelength modes (k small) exhibit positive growth rates (\omega > 0), leading to amplification of fluctuations. The dominant occurs at the maximum growth rate k_{\max} = \sqrt{ - \frac{1}{2\kappa} \frac{\partial^2 f}{\partial \phi^2} }, defining an initial characteristic length scale \lambda \sim 2\pi / k_{\max}. In the early stage, this results in the rapid formation of composition waves or a bicontinuous microstructure with diffuse interfaces. As the decomposition progresses to the late stage, the interconnected domains coarsen to reduce interfacial energy. In the diffusive regime, dominant for solids and viscous fluids, coarsening proceeds via , where smaller domains shrink and dissolve while larger ones grow, governed by the Lifshitz-Slyozov-Wagner (LSW) theory; the average domain radius scales as R \sim t^{1/3}.90054-2) In less viscous fluids, hydrodynamic flow can accelerate coarsening, potentially yielding faster exponents like R \sim t.90128-1) The interfaces remain diffuse throughout, with a mean-field profile \phi(z) \propto \tanh(z / w) across the normal z, where the width w \sim \sqrt{\kappa / |\partial^2 f / \partial \phi^2|} sets the scale over which \phi transitions between the two phases. This tanh form emerges from minimizing the functional for a one-dimensional .

Applications in Physical Systems

Phase Separation in Fluids and Alloys

Phase separation in fluids occurs when a homogeneous of two immiscible liquids demixes into distinct upon changes in or , often exhibiting an (UCST) behavior where cooling below a critical point induces separation. A classic example is the water-phenol system, where the separates into a water-rich and a phenol-rich at temperatures below approximately 66°C, driven by the Flory-Huggins interaction parameter exceeding the critical value for . This demixing process is thermodynamically governed by the landscape, leading to phase domains that grow over time through and hydrodynamic interactions. In the late stages of phase separation in binary fluids, hydrodynamics plays a crucial role in domain coalescence, particularly in bicontinuous morphologies where interconnected s merge via fluid flow. This hydrodynamic coarsening regime results in a linear growth law for the characteristic domain size R \sim t, where t is time, contrasting with the diffusive R \sim t^{1/3} in non-hydrodynamic systems. Simulations and theoretical models confirm that viscous dissipation within boundary layers around domains sustains this linear scaling in the inertial or viscous hydrodynamic limits, enhancing the rate of phase segregation compared to pure . In metallic alloys, phase separation often proceeds in the solid state, where supersaturated solutions decompose into solute-depleted and solute-enriched regions, influencing mechanical properties through microstructural refinement. In steels, such as Fe-Cr alloys, solid-state and lead to phase separation that can form chromium-rich and iron-rich domains, affecting resistance and strength. Similarly, in alloys like III-V compounds (e.g., InGaAs), solid-state immiscibility drives phase separation, resulting in composition modulations that degrade optical and electronic performance if uncontrolled. A key application is , where Guinier-Preston (GP) zones—coherent, solute-rich platelet precipitates—form in aluminum alloys like Al-Cu or Al-Zn-Mg, providing initial strengthening before metastable precipitates evolve. These zones, typically 1-10 nm thick, nucleate via vacancy-assisted during aging, increasing yield strength by impeding motion. Microstructural evolution during phase separation in alloys manifests in distinct reaction types during solidification, such as eutectic and peritectic reactions, which dictate the final morphology. In a eutectic reaction, a of fixed composition transforms simultaneously into two solid phases (e.g., α and β) at the eutectic temperature, yielding a lamellar or rod-like microstructure that enhances in cast irons. In contrast, a peritectic reaction involves a and a primary solid phase reacting to form a new solid phase, often leading to incomplete transformation and pro-eutectic remnants, as seen in peritectic steels where volume contraction can induce cracks. formation, while related to rapid solidification, arises from constitutional undercooling at the solid-liquid interface, producing branched structures that segregate solutes and influence casting homogeneity, though it is mechanistically distinct from diffusional phase separation in the solid state. Industrially, controlling phase separation in alloy casting is essential to minimize defects like , , or cracking, which arise from uneven solidification and solute redistribution. Techniques such as rapid quenching or alloying additions are employed to suppress unwanted precipitation, ensuring uniform microstructures in components like turbine blades. A seminal observation of —characterized by modulated, interconnected structures without barriers—was reported in Cu-Ti s using (TEM), revealing nanoscale Ti-rich and Ti-depleted regions that contribute to age hardening. This 1975 study confirmed spinodal mechanisms in f.c.c. solid solutions, influencing alloy design for high-strength applications. Unique challenges in alloy phase separation include the influence of quenching rates on the decomposition pathway, where faster cooling favors by suppressing barriers and preserving metastable states, while slower rates promote nucleation-and-growth via heterogeneous sites. In duplex stainless steels, for instance, high quenching rates enhance spinodal modulation amplitudes, altering phase fractions and mechanical response. This rate dependence underscores the need for precise thermal processing to tailor microstructures for specific properties.

Phase Separation in Cold Atomic Gases

Phase separation in cold atomic gases occurs in ultracold , such as Bose-Einstein condensates (BECs) and Fermi gases, cooled to temperatures below 1 μK, where quantum coherence dominates and classical is supplemented by quantum effects. In binary BECs composed of two atomic species or hyperfine states, phase separation manifests as a miscible-immiscible transition, where the two components either overlap or spatially separate depending on interspecies interactions. This transition is particularly tunable in these systems due to their dilute nature and precise control over interaction strengths via magnetic Feshbach resonances. Quantum mechanisms driving phase separation in these gases go beyond mean-field approximations, incorporating beyond-mean-field effects through the Lee-Huang-Yang (LHY) correction, which accounts for quantum fluctuations and stabilizes mixtures near the miscibility boundary. In spinor BECs, where atoms have internal spin degrees of freedom, domain formation arises from competing ferromagnetic and antiferromagnetic interactions, leading to spatially separated magnetic phases. These quantum effects distinguish cold atomic gases from classical systems, enabling the study of coherent dynamics and fluctuation-induced phenomena. Experimental milestones include the first observation of controllable phase separation in a dual-species BEC of ^{85}Rb and ^{87}Rb in 2008, where interactions were tuned via a Feshbach resonance to induce immiscibility. A notable example is the 2013 realization of a double-species BEC of ^{23}Na and ^{87}Rb, demonstrating miscible-immiscible transitions through magnetic field tuning of interspecies interactions near a Feshbach resonance at 347.8 G. In spinor BECs, such as F=1 ^{87}Rb, ferromagnetic and antiferromagnetic domains have been visualized, highlighting the role of spin-dependent collisions in phase separation. The dynamics of phase separation in these systems are slow due to the low atomic densities (typically 10^{12}–10^{15} cm^{-3}), allowing of domain growth over milliseconds to seconds. High-resolution imaging via reveals ferromagnetic-antiferromagnetic domains in spinor BECs, with separation proceeding through coherent oscillations before settling into equilibrium configurations. These experiments adapt concepts to quantum fluctuations, where initial instabilities amplify into macroscopic domains. Theoretically, phase separation is described by the coupled Gross-Pitaevskii equations (GPEs) for the two-component wave functions \psi_1 and \psi_2: i \hbar \frac{\partial \psi_1}{\partial t} = \left[ -\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r}) + g_{11} |\psi_1|^2 + g_{12} |\psi_2|^2 \right] \psi_1, i \hbar \frac{\partial \psi_2}{\partial t} = \left[ -\frac{\hbar^2}{2m} \nabla^2 + V(\mathbf{r}) + g_{22} |\psi_2|^2 + g_{12} |\psi_1|^2 \right] \psi_2, where V(\mathbf{r}) is the trapping potential and g_{ij} = 4\pi \hbar^2 a_{ij}/m are interaction strengths with s-wave scattering lengths a_{ij}. Immiscibility occurs when the interspecies coupling satisfies g_{12} > \sqrt{g_{11} g_{22}}, leading to energetic favorability of spatial separation to minimize intercomponent overlap.

Applications in Biological and Soft Matter Systems

Phase Separation in Biological Membranes

Phase separation in biological membranes manifests as the segregation of into distinct domains within the , primarily driven by differences in molecular packing and interactions. In eukaryotic plasma membranes, this process is exemplified by the formation of lipid rafts, which are - and sphingomyelin-rich microdomains existing in a liquid-ordered (Lo) phase. These domains contrast with the surrounding liquid-disordered (Ld) phase, composed mainly of unsaturated phospholipids like , where exhibit higher fluidity and less ordered acyl chain packing. The Lo phase features extended, gel-like acyl chains stabilized by intercalation, while maintaining rapid lateral akin to the Ld phase. The phase behavior of these domains is described by ternary phase diagrams for mixtures such as dioleoylphosphatidylcholine (DOPC), (SM), and , revealing coexistence regions of Lo and Ld phases under physiological conditions. A critical feature is the , where solid-ordered (So), Lo, and Ld phases meet, occurring around 22–37°C depending on composition, encompassing mammalian body temperature and enabling dynamic domain formation . Driving forces for domain stability include line tension at Lo/Ld boundaries, which minimizes boundary length and promotes circular or compact domain shapes, typically on the order of 0.1–10 . Additionally, coupling to —where Lo domains prefer positive —and interactions with proteins, such as GPI-anchored proteins partitioning into Lo regions, further stabilize and refine domain morphology. However, the existence and dynamics of lipid rafts in native cell remain a subject of ongoing and . Experimental evidence for these domains emerged from fluorescence microscopy studies on giant unilamellar vesicles (GUVs), model systems mimicking cell membranes. In 1997, the lipid raft hypothesis was proposed, suggesting these domains as functional platforms for protein organization, based on detergent extraction and biochemical assays showing cholesterol-dependent insolubility of specific lipids and proteins. Direct visualization in GUVs composed of DOPC//cholesterol mixtures demonstrated micron-scale Lo/Ld domains using two-photon fluorescence microscopy with phase-sensitive probes like Laurdan, confirming phase separation at physiological temperatures and revealing domain or tubulation due to line tension. Functionally, lipid rafts serve as sorting platforms for transmembrane proteins, preferentially incorporating (GPI)-anchored proteins and excluding others, facilitating targeted trafficking during or apical sorting in polarized cells. In immune responses, rafts act as signaling hubs, concentrating receptors like the and kinases such as upon , amplifying downstream pathways like MAPK activation in T lymphocytes. This organization enhances signal specificity and efficiency, as demonstrated in studies of formation where rafts cluster to propagate . The presence of domains induces of membrane proteins, characterized by confined motion within Ld or Lo regions interrupted by infrequent hops across boundaries, reducing long-term mobility by factors of 10–100 compared to free . Single-particle tracking in live cells reveals this hop diffusion, with residence times in compartments on the order of milliseconds, attributed to temporary partitioning preferences and barriers at domain edges, impacting processes like receptor clustering and entry.

Liquid-Liquid Phase Separation in Cells

Liquid-liquid phase separation (LLPS) in cells involves the concentration of biomolecules such as proteins and into dynamic, membraneless compartments through weak, multivalent interactions, often mediated by intrinsically disordered regions (IDRs) in proteins. These IDRs enable transient, low-affinity contacts that drive the formation of biomolecular condensates, which behave like liquids and facilitate within the and . A seminal example is the P granules in C. elegans germ cells, which assemble via LLPS and exhibit liquid-like properties such as rapid fusion and dissolution, allowing them to localize to specific cellular regions through controlled phase transitions. Similarly, nucleoli form through LLPS of proteins like fibrillarin and , creating subnuclear compartments essential for . Theoretical models of LLPS in biological systems emphasize the role of sequence features in promoting phase behavior. The sticker-spacer framework describes IDR-driven LLPS as arising from "stickers"—short motifs like aromatic residues that form weak interactions—and "spacers"—flexible linkers that modulate chain and . In the FUS protein, π-cation interactions between residues and in the low-complexity exemplify this, stabilizing multivalent networks that lower the energy barrier for condensate formation while maintaining fluidity. These interactions are tunable by post-translational modifications, such as , which disrupts cation-π bonds and inhibits LLPS. In cellular physiology, LLPS enables compartmentalization for diverse functions, including stress responses and gene regulation. During cellular stress in the 2010s, discoveries linked LLPS to the assembly of granules, dynamic cytoplasmic condensates rich in mRNA and RNA-binding proteins like G3BP1, which sequester translationally stalled transcripts to protect the proteome. Low-complexity domains in proteins such as hnRNPA1 drive this process via phase separation, preventing pathological aggregation under acute . For gene regulation, LLPS at super-enhancers concentrates transcription factors like and , forming phase-separated hubs that enhance promoter looping and boost expression of cell-identity genes. Experimental techniques have illuminated the liquid-like dynamics of these condensates. Fluorescence recovery after photobleaching (FRAP) assays on P granules revealed rapid material exchange with the surrounding nucleoplasm, with half-recovery times on the order of seconds, confirming their fluid behavior distinct from solid aggregates. Optogenetic tools, advanced in the late , enable light-induced control of LLPS by fusing proteins to light-sensitive domains like CRY2, allowing spatiotemporal manipulation of condensate assembly and disassembly in living cells. Aberrant LLPS contributes to neurodegeneration, particularly in amyotrophic lateral sclerosis (ALS), where mutations in RNA-binding proteins like TDP-43 promote excessive phase separation followed by solidification into toxic aggregates. In ALS patients, TDP-43 low-complexity domains exhibit heightened LLPS propensity, leading to cytoplasmic inclusions that impair RNA processing and neuronal function. RNA binding normally suppresses TDP-43 LLPS, but its depletion in disease states shifts the protein toward pathological gel-like states. Additionally, aberrant LLPS has been implicated in oncogenesis through super-enhancer-driven processes as of 2025.

Experimental and Theoretical Methods

Observation Techniques

Phase separation phenomena can be observed and characterized using a variety of experimental techniques that probe structural, dynamical, and morphological features across length scales from nanometers to micrometers. Optical methods, such as , enable real-time visualization of phase domains in by exploiting fluorescence labeling to distinguish separated phases, allowing quantification of droplet size, shape, and coalescence events. Super-resolution techniques like depletion ( extend this resolution to nanoscale domains, revealing sub-diffraction-limited phase-separated regions in lipid membranes and biomolecular condensates with resolutions down to approximately 50 nm. (FCS) complements these imaging approaches by measuring diffusion dynamics within and across phases, providing insights into molecular mobility and phase boundaries through analysis of fluorescence fluctuations. Scattering techniques offer bulk, ensemble-averaged information on domain sizes and evolution without requiring labeling. Small-angle X-ray scattering (SAXS) is widely used to determine domain sizes during phase separation, with scattering intensity profiles yielding characteristic length scales via Porod analysis or fitting to models like the Ornstein-Zernike function. Similarly, small-angle neutron scattering (SANS) provides contrast based on isotopic differences, enabling studies of phase separation in and alloys, such as quantifying through the evolution of the S(k), where peaks indicate characteristic domain spacing. These methods are particularly effective for early-stage detection, as the exhibits a at wavevector k_m corresponding to the dominant domain size, briefly referencing spinodal growth signatures. Electron microscopy techniques provide high-resolution structural details, especially for preserved samples. Cryo-electron microscopy (cryo-EM) captures frozen hydrated states of phase-separated systems, visualizing nanoscale interfaces and morphologies in biological membranes and protein condensates with atomic-level precision in some cases. Cryo-electron extends this to three-dimensional reconstructions, revealing complex architectures in cellular compartments or alloys by tilting samples to generate projection series. Time-resolved studies track the kinetics of phase separation, often using light scattering to monitor growth rates in real time. Time-resolved light scattering measures intensity fluctuations to quantify early-stage coarsening, with growth laws derived from the temporal shift of the peak, typically following power-law scalings like R(t) ~ t^{1/3} for diffusive . These approaches are adaptable to pump-probe setups for ultrafast dynamics in responsive systems. Observing phase separation presents challenges, including artifacts from sample preparation such as fixation-induced coalescence or altered dynamics, which can mimic or obscure true liquid-like behavior; careful controls like live-cell imaging mitigate these. Quantitative metrics, such as interfacial area per volume derived from invariants or segmentations, are essential for comparing domain evolution across techniques but require validation against preparation effects like during purification.

Modeling Approaches

Continuum models provide a foundational framework for simulating phase separation processes involving conserved order parameters, such as composition in binary mixtures. The Cahn-Hilliard equation, derived from a functional that penalizes sharp interfaces, governs the diffusive dynamics of phase separation by minimizing the total through a fourth-order for the order parameter φ. This model captures and coarsening without explicit tracking of interfaces, making it suitable for large-scale simulations of microstructural evolution in alloys and fluids. Extensions incorporating hydrodynamic effects, known as Model H, couple the order parameter evolution to Navier-Stokes equations, accounting for advective and flow-induced changes in fluid mixtures. Phase-field methods build on continuum approaches by representing interfaces as diffuse regions with a continuous order parameter φ varying smoothly from -1 to 1 across phases. The evolution follows the Allen-Cahn or Cahn-Hilliard form, ∂φ/∂t = -Γ δF/δφ, where F is the Ginzburg-Landau functional including bulk, gradient, and potential terms, and Γ is a mobility coefficient. These methods enable predictive simulations of complex microstructures, such as formation during solidification or domain growth in blends, by resolving motion and curvature-driven effects without remeshing. Numerical implementations often use or methods to solve the coupled nonlinear equations efficiently on adaptive grids. Molecular simulations offer mesoscale and atomistic insights into phase separation, bridging continuum models with microscopic interactions. Dissipative particle dynamics () represents groups of molecules as soft, interacting particles with conservative, dissipative, and random forces, enabling simulations of hydrodynamic behavior and phase separation in complex fluids like or block copolymers at coarse-grained resolutions. methods, particularly kinetic variants based on the , sample equilibrium configurations and dynamics of lattice-based alloys, revealing phase diagrams, critical points, and pathways through acceptance criteria or Kawasaki exchanges for conserved dynamics. Advanced techniques enhance modeling fidelity for specific aspects of phase separation. Classical (DFT) computes equilibrium density profiles and interfacial tensions by minimizing a functional over weighted densities, providing accurate predictions of , adsorption, and phase boundaries in inhomogeneous fluids without stochastic sampling. approaches, such as neural networks trained on simulation data, facilitate parameter fitting and surrogate modeling for complex systems, optimizing parameters or predicting phase boundaries in multicomponent mixtures with reduced computational cost. Validation of these models often involves comparing simulated domain growth kinetics to theoretical benchmarks, such as Lifshitz-Slyozov-Wagner (LSW) scaling, where the characteristic domain size R grows as R ~ t^{1/3} during late-stage coarsening due to Ostwald ripening.90054-3) Phase-field and molecular simulations reproduce this exponent across various systems, confirming the dominance of diffusion-limited transport while deviations highlight hydrodynamic or elastic effects.00059-5)

References

  1. [1]
    Phase Separation - an overview | ScienceDirect Topics
    Phase separation is defined as the process where a stabilizer separates from a protein, potentially compromising stability, and can be investigated using ...
  2. [2]
    [PDF] Phase Diagrams and Phase Separation
    Minimum energy for homogeneous single phase, energy Fn. i.e no phase separation occurs in this case. Composition overall determines the state of the mixture.
  3. [3]
    Liquid-liquid phase separation: Fundamental physical principles ...
    Liquid-liquid phase separation (LLPS) is a captivating phenomenon in which a uniform mixture spontaneously divides into two liquid phases with differing ...
  4. [4]
  5. [5]
    [PDF] Droplet Physics and Intracellular Phase Separation
    Dec 7, 2023 · Phase separation is governed by the laws of thermodynamics, reflecting the competition between entropy and energy. Equilibrium between phases is ...
  6. [6]
    Liquid–liquid phase separation in human health and diseases - Nature
    Aug 2, 2021 · Liquid–liquid phase separation (LLPS) represents a vital and ubiquitous phenomenon underlying the formation of membraneless organelles in eukaryotic cells.
  7. [7]
    13.1: The Gibbs Phase Rule for Multicomponent Systems
    ### Summary of Key Concepts from 13.1: The Gibbs Phase Rule for Multicomponent Systems
  8. [8]
    Biomolecular Phase Separation: From Molecular Driving Forces to ...
    Apr 20, 2020 · In this review, we discuss the role that disorder, perturbations to molecular interactions resulting from sequence, posttranslational ...
  9. [9]
    Van der waals equation of state - AIP Publishing
    Nov 1, 2022 · Van der Waals (vdW) invented his equation of state about 150 year ago. This equation described the liquidgas phase transition and predicted an existence of the ...
  10. [10]
    The Gibbs Phase Rule - Oxford Academic
    A van der Waals bond arises when two atoms or molecules are brought towards each other, and the electronic structure of each atom or molecule creates a small, ...Missing: separation | Show results with:separation
  11. [11]
    [PDF] Polymer-Polymer Phase Behavior
    Mixtures of oil and water that normally macroscopically phase separatecan befinely dispersed by the addition of small amounts of surfactant, which can lead ...
  12. [12]
    [PDF] 11.07.05 Free Energy of Multi-phase Solutions at Equilibrium
    Jun 11, 2005 · Common tangents between the free energy curves of different phases occur in regions where 2 phases are in equilibrium.
  13. [13]
    Common Tangent Construction - an overview | ScienceDirect Topics
    The phase diagram, representing the stable or metastable equilibria between the three phases, is determined by the common tangents of these two free-energy ...
  14. [14]
    [PDF] Thermodynamics of mixing
    The phase boundary is determined by the common tangent of the free energy at the compositions o' and o" corresponding to the two equili- brium phases. OAF mix.
  15. [15]
    The lever rule - DoITPoMS
    A tie-line is drawn through the point, and the lever rule is applied to identify the proportions of phases present. Part of a phase diagram. Intersection of the ...
  16. [16]
    Flory-Huggins Theory - an overview | ScienceDirect Topics
    Flory–Huggins theory is defined as a mean-field, lattice model theory that explains the change in Gibbs free energy upon mixing two dissimilar polymers, ...
  17. [17]
    A multi-step nucleation process determines the kinetics of prion-like ...
    Jul 23, 2021 · Spinodal decomposition represents the “speed limit” of the rate at which molecules can diffuse into dense phase droplets. The area in the two- ...Missing: seminal | Show results with:seminal
  18. [18]
    Understanding spinodal and binodal phase transformations in U-50Zr
    However, experimental evidence of spinodal or spinodal-like decomposition in U-Zr alloys has thus far been scarce. Recently, spinodal-like phase decompositions ...Missing: seminal | Show results with:seminal
  19. [19]
    [PDF] 10.626 Lecture Notes, Nucleation and spinodal decomposition
    In this lecture we will study the onset of phase transformation for phases that differ only in their equilibrium composition, keeping in mind that it is.Missing: binodal | Show results with:binodal
  20. [20]
    Observing classical nucleation theory at work by monitoring phase ...
    Dec 3, 2014 · It is widely accepted that many phase transitions do not follow nucleation pathways as envisaged by the classical nucleation theory.
  21. [21]
    [PDF] Review Progress in Ostwald ripening theories and their applications ...
    A major advance in the theory of Ostwald ripening was made in a paper by Lifshitz and Slyozof [5, 6] and followed by a related paper by Wagner [7] (LSW). In.
  22. [22]
    Lifshitz–Slyozov–Wagner theory (Chapter 12) - Dynamics of Self ...
    The Lifshitz–Slyozov–Wagner (LSW) theory of coarsening is based on the assumption that each interface between a minority phase domain and the majority phase ...
  23. [23]
    Nucleation and droplet growth from supersaturated vapor at ...
    Apr 25, 2016 · The homogeneous nucleation in the supersaturated gas is not to a crystal, but to a liquid-like critical nucleus.
  24. [24]
  25. [25]
    Nucleation of New Phases in Alloy: A Long Way to True - Scirp.org.
    In the 60s, the heyday of the theory of spinodal decomposition began again thanks ... (1975) Spinodal Decomposition in Age-Hardening Cu-Ti Alloys. Acta ...Missing: 1960 | Show results with:1960
  26. [26]
    Hydrodynamic Coarsening of Binary Fluids
    Apr 10, 2000 · Hydrodynamic Coarsening of Binary Fluids. Francisco J. Solis and Monica Olvera de la Cruz. Phys. Rev. Lett. 84, 3350 – Published 10 April 2000.Missing: R ~ | Show results with:R ~
  27. [27]
    Hydrodynamic effects in kinetics of phase separation in binary fluids
    Apr 21, 2023 · Via hydrodynamics-preserving molecular dynamics simulations we study growth phenomena in a phase-separating symmetric binary mixture model.Missing: phenol | Show results with:phenol
  28. [28]
    Segregation, precipitation, and phase separation in Fe-Cr alloys
    Dec 28, 2015 · In this work, we investigate segregation, precipitation, and phase separation in Fe-Cr systems analyzing the physical mechanisms behind the observed phenomena.
  29. [29]
    The Thermodynamics and Kinetics of Phase Separation in III-V ...
    Mar 30, 2024 · In order to understand the role of surface diffusion in phase separation in compound semiconductor alloys, it is useful to study the system in ...
  30. [30]
    Solid solution decomposition and Guinier-Preston zone formation in ...
    Mar 21, 2018 · We show that the decomposition of the solid solution forming platelets of copper, known as Guinier-Preston (GP) zones, includes several stages.Missing: steels semiconductors
  31. [31]
    [PDF] EUTECTICS, PERITECTICS AND MICROSTRUCTURE SELECTION
    Eutectics are invariant points where liquid transforms directly into two solid phases, requiring a second phase to nucleate and grow.
  32. [32]
    [PDF] Metal Casting Defects and Prevention - iaeme
    The proper classification and identification of a particular defect is the basic need to correct and control the quality of casting. Most of these castings are ...
  33. [33]
    Influence of the Quenching Rate on the Spinodal Decomposition in ...
    Aug 6, 2025 · Steels of equivalent composition after undergoing the seemingly same heat treatment exhibit different microstructural and mechanical evolutions.
  34. [34]
    [PDF] 5 Homogeneous Second-Phase Precipitation
    Decomposition Mechanisms: Nucleation and Growth versus Spinodal ... Decomposition Kinetics in Alloys Pre-Decomposed During Quenching . . 391. 5.7.6.
  35. [35]
    Lee-Huang-Yang effects in the ultracold mixture of 2 3 N a and 8 7 R ...
    Sep 14, 2021 · In this work, we study in two different ways the LHY effects in a double BEC of Na 23 and Rb 87 atoms with tunable attractive interspecies ...
  36. [36]
    Phase separation and dynamics of two-component Bose-Einstein ...
    Jul 11, 2016 · Two-component Bose-Einstein condensates enable the investigation of this scenario in a particularly well controlled setting.
  37. [37]
    Functional rafts in cell membranes - Nature
    Jun 5, 1997 · Simons, K., Ikonen, E. Functional rafts in cell membranes. Nature 387, 569–572 (1997). https://doi.org/10.1038/42408. Download citation. Issue ...
  38. [38]
    Sphingomyelin/Phosphatidylcholine/Cholesterol Phase Diagram - NIH
    The ternary system palmitoylsphingomyelin (PSM)/palmitoyloleoylphosphatidylcholine (POPC)/cholesterol is used to model lipid rafts.
  39. [39]
    Line Tension at Fluid Membrane Domain Boundaries Measured by ...
    May 16, 2007 · Line tension at the domain phase boundary controls the kinetics of phase separation and domain sizes. In vivo, line tension could be an ...
  40. [40]
    The role of lipid rafts in signalling and membrane trafficking in T ...
    Nov 15, 2001 · Upon triggering, lipid rafts concentrate in the immunological synapse, gathering together specific membrane proteins. Lck becomes activated and ...
  41. [41]
    Phospholipids undergo hop diffusion in compartmentalized cell ...
    The diffusion rate of lipids in the cell membrane is reduced by a factor of 5–100 from that in artificial bilayers. This slowing mechanism has puzzled cell ...
  42. [42]
    Rapid Hop Diffusion of a G-Protein-Coupled Receptor in the Plasma ...
    In both models, membrane proteins and lipids can hop from a compartment to an adjacent one, probably when thermal fluctuations of the membrane and the membrane ...μor Diffusion Observed By... · Comparison Of μor's... · μor Diffusion Data After...
  43. [43]
    Germline P Granules Are Liquid Droplets That Localize by ... - Science
    Here we show that P granules exhibit liquid-like behaviors, including fusion, dripping, and wetting, which we used to estimate their viscosity and surface ...
  44. [44]
    FUS Phase Separation Is Modulated by a Molecular Chaperone and ...
    Apr 19, 2018 · Here, we report that cooperative cation-π interactions between tyrosines in the LC domain and arginines in structured C-terminal domains also ...
  45. [45]
    Coactivator condensation at super-enhancers links phase ... - Science
    Super-enhancers are clusters of enhancers bound by master transcription factors that concentrate high densities of coactivators and the transcription apparatus ...Coactivator Condensation At... · Brd4 And Med1 Coactivators... · Unsupported
  46. [46]
    Detecting and quantifying liquid–liquid phase separation in living ...
    Dec 16, 2022 · Cells contain numerous substructures that have been proposed to form via liquid–liquid phase separation (LLPS). ... Confocal microscopy · Imaging ...Introduction · Results and discussion · Methods · References
  47. [47]
    STED microscopy detects and quantifies liquid phase separation in ...
    STED microscopy detects and quantifies liquid phase separation in lipid membranes using a new far-red emitting fluorescent phosphoglycerolipid analogue · Authors.Missing: nanoscale | Show results with:nanoscale
  48. [48]
    Fluorescence Correlation Spectroscopy and Phase Separation
    Fluorescence Correlation Spectroscopy (FCS) provides a versatile approach to estimate phase boundaries of single-component and multicomponent solutions.
  49. [49]
    Small-angle x-ray-scattering study of phase separation and ...
    May 1, 1999 · We report on a small-angle x-ray-scattering (SAXS) and differential scanning calorimetry study of phase separation and crystallization in ...
  50. [50]
    Small-angle neutron scattering quantification of phase separation ...
    Small-angle neutron scattering (SANS) was applied to quantify the nanostructural evolution during spinodal decomposition in a 25Cr-7Ni (wt.%) super duplex ...
  51. [51]
    Small-angle X-ray scattering from the concentrated bulk phase ...
    Jan 18, 2017 · SAXS measurements were performed on the macroscopically phase-separated concentrated phase of an aqueous solution of the thermosensitive ...
  52. [52]
    Cryo-EM images of phase-separated lipid bilayer vesicles ... - PubMed
    Sep 3, 2024 · Using such models, we recently showed that cryogenic electron microscopy (cryo-EM) can detect phase separation in lipid vesicles based on ...
  53. [53]
    Quantitative spatial analysis of chromatin biomolecular condensates ...
    Cryoelectron tomography (cryo-ET) is a potentially powerful technique for structural analyses of condensates because it enables visualization of all individual ...
  54. [54]
    Time-resolved light scattering studies on kinetics of phase ...
    Time-resolved light scattering studies on kinetics of phase separation and phase dissolution of polymer blends. 1. Kinetics of phase separation of a binary ...
  55. [55]
    Fixation can change the appearance of phase separation in living cells
    Nov 29, 2022 · Our work reveals that PFA fixation changes the appearance of LLPS from living cells, presents a caveat in studying LLPS using fixation-based methods,
  56. [56]
    Considerations and challenges in studying liquid-liquid phase ...
    In cells, excess protein may be stored in membraneless organelles and will enter the dilute phase as needed when protein levels drop. LLPS can locally ...