Fact-checked by Grok 2 weeks ago

Miscibility gap

A miscibility gap is a region in the of a or multicomponent where the components exhibit limited , leading to into two or more distinct phases across a specific range of compositions and temperatures. This separation occurs because the of the homogeneous exceeds that of the coexisting phases, rendering mixing thermodynamically unfavorable. The thermodynamic foundation of a miscibility gap is rooted in the of mixing, often modeled using the regular solution approximation, which accounts for non-ideal contributions while assuming ideal . In this model, the interaction parameter Ω quantifies the energetic penalty for unlike-pair interactions; when Ω is sufficiently positive and temperature-dependent, the curve develops a double-well shape, with common tangent constructions defining the phase boundaries. The gap typically narrows or closes at higher temperatures near a , where thermal overcomes enthalpic repulsion, allowing complete . Miscibility gaps are prevalent in diverse systems, including metallic alloys such as Pb-Sn and Ag-Cu, where they manifest as two-phase regions (e.g., α and β solids) below the liquidus, influencing solidification and mechanical properties. In liquid systems like melts, gaps appear in compositions with low alumina content (e.g., <2 wt% Al₂O₃ in Na₂O-SiO₂), driven by polymerization differences and cation associations, and can be suppressed by additives or pressure. Within the gap, metastable regions enable spinodal decomposition, a diffusion-driven mechanism forming modulated microstructures without nucleation barriers, which is critical for applications in semiconductors and glasses.

Fundamentals

Definition

A miscibility gap is a region in the phase diagram of a mixture where two or more components exhibit immiscibility, resulting in phase separation into distinct phases, such as two coexisting liquid phases or two solid phases. This occurs when the mixture's composition and conditions, like temperature, fall within boundaries where the components cannot form a single homogeneous phase, leading to spontaneous demixing. In mixtures, complete miscibility implies that components can dissolve in each other across all proportions under given conditions, forming a single phase, whereas partial miscibility allows solubility only up to certain limits, beyond which a miscibility gap emerges. The gap specifically denotes the concentration and temperature range where these solubility limits are exceeded, causing the system to split into multiple phases with compositions at the gap's boundaries. For instance, in binary mixtures, this gap appears as a lens-shaped area in temperature-composition diagrams, separating regions of single-phase stability from two-phase coexistence. The concept of miscibility gaps was first described in the context of binary mixtures during the 19th century, with observations of phase separation in liquid systems, and was formalized through J. Willard Gibbs' phase rule in the 1870s, which provided the thermodynamic framework for understanding equilibrium in multi-phase systems. Gibbs' work integrated composition variables into phase equilibria, enabling precise prediction of conditions for immiscibility. Basic prerequisites for a miscibility gap include differences in intermolecular forces between unlike components compared to like components, which contribute to a positive enthalpy of mixing that can outweigh the stabilizing entropy of mixing. This imbalance in interaction energies leads to thermodynamic instability in certain composition ranges, promoting phase separation as the system minimizes its free energy.

Thermodynamic Principles

The formation of a miscibility gap in binary mixtures is governed by the Gibbs free energy of mixing, \Delta G_{\text{mix}}, which determines the thermodynamic stability of homogeneous phases. For an ideal solution, \Delta G_{\text{mix}} = \Delta H_{\text{mix}} - T \Delta S_{\text{mix}}, where \Delta H_{\text{mix}} is the enthalpy of mixing, T is the temperature, and \Delta S_{\text{mix}} is the entropy of mixing. In non-ideal systems, such as those exhibiting limited solubility, \Delta G_{\text{mix}} often features a double-well shape or a positive maximum as a function of composition, indicating instability and a tendency for phase separation into two coexisting phases with lower overall free energy. This instability arises when the second derivative \partial^2 \Delta G_{\text{mix}} / \partial x^2 < 0 over a composition range x, defining the spinodal region within the broader miscibility gap. At equilibrium, the compositions of the coexisting phases are identified using the common tangent construction on the \Delta G_{\text{mix}} versus composition curve. This graphical method draws a tangent line that touches the free energy curve at the compositions of the two phases, ensuring equal chemical potentials and minimizing the total free energy of the system. The relative amounts of each phase follow the lever rule, where the fraction of one phase is proportional to the distance from the overall composition to the tangent points. The driving force for immiscibility primarily stems from the enthalpy term: positive \Delta H_{\text{mix}} reflects unfavorable intermolecular interactions between unlike components, such as in regular solution models where \Delta H_{\text{mix}} = \Omega x_A x_B and \Omega > 0. In contrast, the entropic contribution favors mixing, given by the ideal \Delta S_{\text{mix}} = -R \sum_i x_i \ln x_i (per mole), which is always positive for $0 < x_i < 1 and promotes homogeneity through increased configurational disorder. The competition between these terms results in a miscibility gap that typically widens at lower temperatures, where the -T \Delta S_{\text{mix}} term diminishes relative to \Delta H_{\text{mix}}. The boundaries of the miscibility gap are delimited by critical solution temperatures, known as the upper critical solution temperature (UCST) or lower critical solution temperature (LCST), where the gap closes and complete miscibility is achieved across all compositions. Below the UCST (common in systems with positive \Delta H_{\text{mix}}), or above the LCST (seen in hydrogen-bonding systems with negative \Delta H_{\text{mix}}), the two-phase region exists due to the curvature of \Delta G_{\text{mix}} at the critical point, where both the first and second derivatives with respect to composition vanish. For polymeric systems, the Flory-Huggins theory provides a lattice-based model for \Delta G_{\text{mix}}, incorporating the interaction parameter \chi to quantify enthalpic interactions. For polymer solutions (solvent with N=1, polymer with large N), it approximates \Delta G_{\text{mix}} / RT = (1-\phi) \ln(1-\phi) + \phi \ln \phi + \chi \phi (1-\phi), where \phi is the volume fraction of polymer, with immiscibility when \chi > 0.5. For symmetric polymer blends (both components polymers of equal degree of polymerization N), the entropy terms are reduced: \Delta G_{\text{mix}} / RT = (\phi / N) \ln \phi + ((1-\phi)/N) \ln (1-\phi) + \chi \phi (1-\phi), leading to phase separation when \chi > 2/N, a much smaller threshold for large N due to diminished entropic stabilization from chain connectivity. This mean-field approximation highlights how polymeric architecture promotes immiscibility compared to small-molecule mixtures.

Representation in Phase Diagrams

Binary Systems

In binary systems, miscibility gaps are represented in isobaric temperature-composition (T-x) phase diagrams at constant pressure, where the gap manifests as a lens-shaped two-phase region bounded by solvus lines that define the solubility limits of each component in the coexisting phases. These diagrams illustrate how, within the gap, mixtures separate into two distinct phases with differing compositions, while outside the gap, the system remains fully miscible as a single phase. The solvus lines converge at a critical point, marking the boundary beyond which complete miscibility is achieved. Miscibility gaps in binary systems can exhibit either (UCST) or (LCST) behavior. In UCST systems, the gap exists below the critical temperature, narrowing as temperature increases until the solvus lines meet at the UCST, above which the phases become fully miscible; a representative example is the methyl acetate-carbon disulfide mixture, where the UCST occurs at 39°C. Conversely, LCST systems display the gap above the critical temperature, with immiscibility increasing as temperature rises and the solvus lines converging at the LCST from below; poly(N-isopropylacrylamide)- is a classic example, exhibiting an LCST near 32°C due to polymer coil collapse. Within the miscibility gap, tie lines are horizontal constructs in diagrams that connect the compositions of the two coexisting at a fixed , allowing determination of phase equilibria. The relative proportions of these phases for an overall falling in the gap are calculated using the : the fraction of the phase with composition x_\alpha is (x - x_\beta)/(x_\alpha - x_\beta), where x is the overall and x_\alpha, x_\beta are the endpoint compositions along the tie line, analogous to balancing a . In isothermal pressure-composition (P-x) diagrams, the miscibility gap's extent depends on , often narrowing or closing with increasing due to enhanced molecular interactions that favor mixing. This sensitivity highlights how compressive forces alter phase boundaries in mixtures.

Multicomponent Systems

In multicomponent systems with three or more components, miscibility gaps extend beyond the simplicity of representations, introducing additional that result in more complex behaviors. phase diagrams, which depict systems with three components, are typically projected onto equilateral triangular plots, often combined with as a vertical axis to form triangular T-x-y diagrams. In these diagrams, miscibility gaps manifest as polyhedral regions bounded by surfaces where occurs, such as lens-shaped or irregular volumes isolating immiscible liquid or solid s. Liquidus surfaces in ternary diagrams represent the boundary above which a single liquid phase exists, sloping downward from the melting points of pure components and often forming valleys or ridges separated by invariant lines like eutectic or peritectic reactions. Solvus surfaces, analogous to those in binaries, delineate the limits of solid solution stability, enclosing regions where solid phases separate into multiple compositions due to limited solubility. These surfaces in multicomponent diagrams can intersect to create multifaceted polyhedra, complicating the visualization of phase equilibria compared to two-dimensional binary plots. Isothermal sections of ternary phase diagrams, sliced at constant temperature, provide planar views of phase assemblages across the composition triangle. Within these sections, miscibility gaps appear as areas separated by tie lines connecting coexisting two-phase regions, while three-phase coexistence is indicated by tie triangles—enclosed areas where the vertices represent the compositions of the three equilibrium phases. Such tie triangles highlight invariant equilibria under the Gibbs phase rule for ternary systems at fixed temperature and pressure, enabling the determination of phase fractions via the lever rule extended to triangular geometry. Mapping gaps in multicomponent systems poses significant challenges due to the high dimensionality, requiring extensive experimental data for validation and often leading to incomplete diagrams for or higher-order systems. Computational tools like the (Calculation of Phase Diagrams) method address these by minimizing the across phases using thermodynamic models extrapolated from lower-order subsystems, accurately predicting gap boundaries without initial composition guesses. This approach has been refined through algorithms that ensure convergence for arbitrary component numbers, facilitating reliable predictions of polyhedral miscibility regions.

Phase Separation Mechanisms

Binodal and Spinodal Decomposition

The binodal curve defines the boundary of the miscibility gap in a phase diagram, marking the compositions at which two phases coexist in equilibrium. It is determined by the common tangent construction applied to the Gibbs free energy of mixing as a function of composition, where the tangent lines connect points of equal chemical potential between the coexisting phases, ensuring the overall free energy is minimized. Within the miscibility gap, the spinodal curve delineates the boundary between the metastable and unstable regions, specifically where the second derivative of the free energy of mixing with respect to composition, \partial^2 \Delta G_{\text{mix}} / \partial c^2, equals zero. This condition signifies the onset of diffusive instability, as small composition fluctuations become amplified rather than damped, leading to spontaneous phase separation without an energy barrier. Spinodal decomposition occurs in the unstable region inside the spinodal curve, where infinitesimal composition fluctuations grow exponentially through diffusion, resulting in the formation of interconnected, modulated domains. This process is mathematically described by the Cahn-Hilliard equation, \frac{\partial c}{\partial t} = \nabla \cdot \left( M \nabla \frac{\delta F}{\delta c} \right), where c is the composition, M is the , and F is the functional incorporating both bulk and gradient contributions. Unlike and growth, which requires overcoming a free energy barrier to form distinct droplets, is barrierless and initially faster, driven solely by thermodynamic instability. Early stages of spinodal decomposition are experimentally observed using small-angle X-ray scattering (SAXS), which detects the characteristic scattering peaks corresponding to the wavelength of composition modulations as they evolve over time. SAXS profiles typically show an initial peak shift to higher wavevectors followed by growth in intensity, confirming the diffusive amplification predicted by .

Nucleation and Growth

In the metastable region of a miscibility gap, located between the and spinodal curves, phase separation proceeds via and growth mechanisms, where an energy barrier must be overcome to form stable domains of the minority phase. This contrasts with the barrierless spinodal decomposition in the unstable region inside the spinodal. Nucleation involves the formation of small, critical-sized clusters that serve as precursors to the separated phases, driven by the competition between the bulk gain and the interfacial energy penalty. Classical nucleation theory describes this process by modeling the free energy change for forming a spherical of r as \Delta G(r) = \frac{4}{3} \pi r^3 \Delta G_v + 4 \pi r^2 \sigma, where \Delta G_v is the volumetric difference between the parent and nucleating phases (negative in the metastable region), and \sigma is the interfacial . The critical size r^* occurs at the maximum of this function, r^* = -\frac{2\sigma}{\Delta G_v}, yielding a barrier \Delta G^* = \frac{16\pi \sigma^3}{3 (\Delta G_v)^2}. The rate is then I = I_0 \exp\left( -\frac{\Delta G^*}{k_B T} \right), where I_0 is a kinetic prefactor, k_B is Boltzmann's constant, and T is temperature; this barrier height decreases with increasing undercooling or deviation from the , facilitating . Nucleation can be homogeneous, occurring uniformly within the bulk without preferential sites, or heterogeneous, catalyzed by impurities, walls, or existing phases that lower the effective \Delta [G](/page/G)^* by reducing the interfacial term. In immiscible alloys like undercooled Cu-Co, homogeneous dominates for the Co-rich when the Cu-rich envelops potential heterogeneous sites, suppressing surface-catalyzed events and leading to a higher of approximately 67 k_B T. Heterogeneous prevails closer to the , where the driving force is weaker but sites are more effective. Following nucleation, the growth stage involves diffusion-limited coarsening of the domains, where smaller particles dissolve and larger ones grow to minimize total interfacial energy, as described by Lifshitz-Slyozov-Wagner (LSW) theory. In this regime, the average domain radius R scales as R \sim t^{1/3}, with the constant of proportionality increasing with the volume fraction of the minority phase; for dilute systems, \bar{R}^3(t) = \bar{R}^3(0) + Kt, where K depends on and . This t^{1/3} kinetics arises from the balance of solute fields around particles, assuming no interactions or coalescence. Several factors influence the nucleation and growth in the metastable region, including undercooling, which amplifies |\Delta G_v| and reduces \Delta G^*; composition deviations from the , which modulate ; and additives or impurities that act as heterogeneous sites or alter interfacial tension \sigma. For instance, in III-V alloys like InGaAs, higher group V concentrations narrow the miscibility gap and suppress barriers by up to 64°C in critical , while differences between phases can fully eliminate the gap during growth. In electrochemical systems, additives like reduce and promote by narrowing the metastable zone width.

Examples

Metallic Alloys

Miscibility gaps are prevalent in substitutional metallic alloys, such as the Cu-Ni system, where they manifest as regions of phase immiscibility in the solid state due to thermodynamic instabilities that drive atomic ordering and subsequent . In Cu-Ni alloys, the miscibility gap appears below approximately 600 K, with phase boundaries at around 34 at.% Cu and 74 at.% Cu at 573 K, leading to or nucleation-driven separation that results in nanoscale compositional modulations. Similarly, in permanent magnet alloys (e.g., Fe-Ni-Al-Co systems), the miscibility gap in the Fe-Ni-Al subsystem exhibits an asymmetric form, with the α and α′ phases separating below the , influencing structures through controlled phase partitioning. These gaps play a critical role in age-hardening processes within metallic alloys, where supersaturated solid solutions are formed at high temperatures and then quenched to trap solutes, followed by controlled aging to induce within the gap region. This exploits the thermodynamic drive for , forming coherent or semi-coherent precipitates that impede motion and enhance strength, as seen in aluminum alloys where Guinier-Preston zones emerge within the metastable miscibility gap during early aging stages. In systems like Au-Ni, the gap enables similar hardening by limiting solute at lower temperatures, resulting in ordered phases that increase hardness without significant loss of . Recent advancements leverage these gaps for microstructural engineering in additively manufactured alloys. A 2025 study on laser-processed -8H demonstrated that low-temperature annealing (e.g., 500–600°C) exploits the solid-state miscibility gap to refine nanoscale α₁ and α₂ phases, achieving enhanced (up to 1.5 kOe) and while maintaining high-temperature stability, thus improving hard magnetic performance over traditional cast . In systems, such as Ag-Cu-Se, liquid miscibility gaps identified in 2025 analyses reveal two extensive immiscible regions, promoting the formation of core-shell microstructures during solidification, where Se-rich cores are encapsulated by Ag-Cu shells, enabling tailored morphologies for electronic applications. The presence of miscibility gaps profoundly affects mechanical properties in metallic alloys, often providing strengthening through fine-scale but risking embrittlement if uncontrolled. Precipitation within the gap, as in age-hardened alloys, can increase strength by 50–200% via obstacle formation to dislocations, yet excessive separation in systems like Fe-Cr leads to the , where α′ ferrite formation depletes Cr from the matrix, reducing toughness by up to 80% and promoting . Alloying elements like or can modulate the gap width to balance these effects, narrowing it to favor strengthening over in duplex stainless steels.

Mineral and Geological Systems

In mineral and geological systems, miscibility gaps are prominent in the solid solution series, where they manifest as regions of compositional instability leading to exsolution textures during cooling. The Huttenlocher gap occurs between approximately An55 and An95 (where An denotes the mole fraction), resulting in high-temperature exsolution of sodic and calcic phases in intermediate to calcic . Similarly, the Bøggild gap spans An39-An48 and An53-An63, producing lamellar intergrowths of and compositions that exhibit due to optical . The peristerite gap, found around An5-An15, involves fine-scale exsolution between albite-rich and phases, often observed in low-grade metamorphic or slowly cooled igneous rocks. These miscibility gaps have significant geological implications, particularly in igneous rocks where exsolution textures form diagnostic features for thermometry and cooling history . Lamellae spacing in peristerite or Bøggild intergrowths, for instance, records cooling rates from hundreds to thousands of degrees per million years, providing insights into plutonic emplacement and thermal evolution. Such textures arise via slow cooling through the solvus surface, promoting diffusional unmixing and resulting in perthitic intergrowths that enhance mineral stability under subsolidus conditions.

Liquid and Polymeric Mixtures

Miscibility gaps in liquid mixtures manifest as regions where two fluids do not fully mix, leading to under certain conditions. A classic example is the , where immiscibility arises due to the low of non-polar oils in polar , resulting in distinct phases. This phenomenon is particularly evident in processes, where "oiling out" occurs as a liquid-liquid (LLPS) driven by a miscibility gap in the solute-solvent system. During cooling or antisolvent addition, the solute-rich phase forms an oily droplet dispersion instead of solid crystals, complicating product isolation and purity control. In some aqueous alcohol systems, such as -water mixtures, temperature-dependent behavior can influence effective , particularly when combined with solutes exhibiting (LCST) transitions. LCST denotes the temperature above which occurs in otherwise miscible mixtures, often due to entropy-driven changes in bonding and hydrophobic interactions. For instance, adding to water-based solutions of thermoresponsive polymers lowers the LCST, enhancing for applications like controlled release, though pure -water remains fully miscible across typical ranges. Polymeric mixtures often display pronounced gaps due to the entropic penalties of mixing long chains and unfavorable enthalpic interactions. In blends like () and (), the Flory-Huggins interaction parameter χ exceeds the critical value of approximately 2 (scaled by ), promoting immiscibility and microphase separation into domains. This leads to morphologies such as lamellae or cylinders, which influence mechanical properties like in rubber-toughened plastics. The χ parameter, quantifying pairwise interactions, highlights how polar differences between and drive domain formation on the nanoscale. Spinodal decomposition in liquid mixtures occurs rapidly within the unstable region of a miscibility gap, as seen in partially miscible fluids like metallic alloys in the melt state. Upon into the , composition fluctuations amplify spontaneously, forming interconnected bicontinuous structures without barriers. This diffusion-driven process contrasts with in metastable regions and is modeled via Cahn-Hilliard equations, yielding scales on the order of micrometers in fluids with moderate viscosities. Recent advancements in sustainable chemistry have emphasized miscibility data for green solvents to optimize eco-friendly processes. A 2025 study compiled miscibility tables for 28 green solvents, including bio-based options like and cyrene, revealing patterns such as full miscibility among polar aprotic alternatives but gaps with non-polar hydrocarbons. These tables guide solvent selection in extraction and reaction media, reducing environmental impact by favoring recyclable, low-toxicity pairs over volatile organic compounds. Miscibility gaps in fluids also induce hydrodynamic instabilities, notably Rayleigh-Taylor and Kelvin-Helmholtz types, as explored in 2025 investigations. The Rayleigh-Taylor instability arises when a denser overlays a less dense one under , accelerated by phase separation in binary mixtures with temperature-sensitive gaps, leading to patterns. Similarly, Kelvin-Helmholtz instability develops at interfaces in miscible-to-immiscible transitions, promoting vortex formation and enhanced mixing in flows. Phase-field simulations of these phenomena in partially miscible binary fluids demonstrate how gap width influences instability growth rates, with applications in and stability.

Applications and Developments

Thermal Energy Storage

Miscibility gap alloys (MGAs) serve as innovative phase change materials (PCMs) in thermal energy storage systems, leveraging the latent heat of fusion in binary metal systems exhibiting a miscibility gap. These alloys, such as Ga-In or Bi-Sn, feature inverse microstructures where a low-melting fusible phase (e.g., liquid In or Sn) is encapsulated within a solid skeletal matrix of the higher-melting component (e.g., solid Ga or Bi), enabling operation through the miscibility gap during melting and solidification. This design provides high latent heat capacities, typically ranging from 180 to 200 kJ/kg for systems like Fe-Cu or Bi-Sn, surpassing many conventional organic or salt-based PCMs while maintaining structural integrity. The primary advantages of MGAs in thermal storage include exceptional shape stability, as the solid matrix prevents volume expansion or leakage during phase transitions, high thermal conductivity (often 50-200 times greater than salt hydrates or paraffins due to their metallic nature, typically 20-100 W/m·K), and compatibility with electro- energy storage (ETES) systems for efficient . These properties allow MGAs to address key limitations of traditional PCMs, such as low and container corrosion, making them suitable for applications requiring reliable, long-duration storage without encapsulation needs. For instance, in (CSP) systems, conceptual designs integrate MGAs for high-temperature storage (up to 500°C), with numerical analyses demonstrating reduced sizes and improved efficiency through enhanced rates. Recent advancements highlight MGAs' practical deployment, including the 2025 reboot and commissioning of a 5 MWh by MGA Thermal for green and thermal storage, enabling continuous renewable production at costs competitive with fossil fuels. Additionally, a 2025 from the of Newcastle explored novel MGA compositions for storage, focusing on high-temperature variants (e.g., Al-Si or Fe-based systems) to meet demands in CSP and , with prototypes showing over 1,000 cycles at 400-600°C. These developments underscore MGAs' potential for scalable ETES, such as modular blocks stackable up to multi-MWh capacities. Despite these benefits, MGAs face challenges including high material costs due to rare metals like gallium and indium (estimated at $10-50/kWh for systems), which can limit commercialization, and concerns over long-term cycling stability, where repeated thermal exposure may degrade the microstructure after 500-1,000 cycles in some alloys. Ongoing research addresses these through alloy optimization and cost modeling to enhance economic viability for widespread adoption in renewable energy integration.

Battery and Electrochemical Materials

In lithium-ion battery cathodes, miscibility gaps play a critical role in phase stability and performance degradation, particularly in high-voltage spinel materials like LiNi0.5Mn1.5O4 (LNMO). The ordered spinel structure of LNMO exhibits a room-temperature miscibility gap during lithium extraction, leading to two-phase coexistence between Li-rich and Li-poor phases, which induces volume changes and mechanical stress that accelerate capacity fade over cycling. Recent advancements involve multi-element doping with Si, Ti, and Ge to tailor this miscibility gap, suppressing phase separation by stabilizing the solid solution phase and enhancing charge/discharge reversibility, resulting in improved cycling stability up to 1000 cycles at 4.75 V. Cation mixing further influences miscibility gaps in these cathodes, where Ni/Mn disorder narrows the gap under high current densities, promoting a more continuous reaction over abrupt two-phase transitions. Studies from 2022 demonstrate that increasing current density from 0.1C to 5C reduces the miscibility gap width in LNMO variants, enhancing rate capability but risking incomplete phase reversion and long-term instability if not managed. This narrowing effect stabilizes the structure during fast charging, crucial for applications, though excessive mixing can lead to irreversible capacity loss. In electrolytes, miscibility gaps in mixtures enable stimuli-responsive behaviors, such as temperature-triggered for enhanced safety. For instance, poly(ethylene oxide)/ systems exhibit (LCST) behavior, where heating above ~60°C induces into polymer-rich and -rich phases, rapidly increasing resistance to shut down and prevent in lithium-ion batteries. This reversible separation maintains electrochemical performance at ambient temperatures while providing overcharge protection, with prototypes showing negligible capacity loss after multiple thermal cycles. The implications of gaps in these materials include accelerated fade due to and associated side reactions, such as manganese dissolution in LNMO cathodes, which reduces active material utilization by up to 20% after 500 cycles. Strategies like post-synthesis annealing at 700–900°C control cation ordering and narrow the gap, minimizing by promoting homogeneous solid solutions and extending cycle life. In emerging applications, overcoming the gap in / systems via plasma-assisted (PAMBE) enables cubic InxGa1-xN alloys across the full composition range, supporting high-efficiency semiconductors for in next-generation batteries with reduced polarization losses.

Nanomaterials and Emerging Uses

Miscibility gaps play a pivotal role in the design of multicomponent nanoparticles, enabling the of complex nanostructures such as core-shell particles by predicting immiscible boundaries in temperature-composition space. A 2022 review highlights how thermodynamic modeling of these gaps guides precursor ratios to achieve stable, phase-separated architectures in , expanding the compositional space beyond traditional systems. This approach facilitates tailored properties like enhanced catalytic activity and magnetic responsiveness at the nanoscale. In binary fluids exhibiting temperature-sensitive miscibility gaps, hydrodynamic instabilities offer innovative pathways for nanoscale applications. The Rayleigh-Taylor instability, where a denser fluid overlies a lighter one under acceleration, has been modeled using phase-field methods to capture transitions from immiscible to miscible states, revealing distinct growth regimes influenced by interfacial tension and diffusion. Such dynamics are leveraged for , where stimuli-responsive enables precise release mechanisms in biological environments. Similarly, the Kelvin-Helmholtz instability, driven by velocity shear across interfaces, promotes mixing in partially miscible fluids, with simulations showing suppressed wave growth as miscibility increases, which is critical for microfluidic devices in technologies. Advancements in exploit gaps to extend compositions. In () growth of cubic (In,Ga)N films, strain management overcomes the GaN/InN gap, achieving indium contents up to 40% without , as demonstrated on GaN/AlN templates. This enables polarization-free optoelectronic devices with tunable bandgaps for UV-visible applications. has emerged as a tool for predicting gaps in complex systems, particularly body-centered cubic (BCC) . models trained on data accurately forecast phase diagrams, identifying stable regions and gaps with errors below 5% for and BCC compositions, accelerating materials discovery for high-temperature applications. Ab initio calculations combined with Bayesian learning have clarified discrepancies in the Ti-V phase diagram, attributing the observed miscibility gap in the β-phase to oxygen impurities that stabilize clustering at concentrations as low as 1 at.%. These insights refine design for components, confirming no intrinsic gap in oxygen-free systems. In , updated miscibility tables for 28 bio-based and sustainable solvents address gaps in traditional datasets, evaluating pairwise compatibilities to minimize hazardous mixtures in . This resource supports scalable, eco-friendly processes by predicting phase behavior for solvent selection in pharmaceutical and production.

References

  1. [1]
    [PDF] Lecture 18: 11.09.05 Binary systems: miscibility gaps and eutectics
    Sep 11, 2005 · Key points: 1. Common tangents to free energy curves define composition regions where phase separation (two- phase equilibria) occurs.
  2. [2]
  3. [3]
  4. [4]
    Miscibility Gap - an overview | ScienceDirect Topics
    For a reference miscibility gap, the excess Gibbs energy of a phase (it does not matter whether it is a liquid or a solid) was taken to be of the form Gex = 20 ...
  5. [5]
    Historic Note No. 1: Gibbs' Phase Rule - Thermo-Calc Software
    Gibbs' phase rule, released in 1875, included composition in thermodynamics, but was not recognized until 1901 by Bakhuis Roozeboom.
  6. [6]
    Gibbs' Phase Rule: Where it all Begins - SERC (Carleton)
    Aug 8, 2007 · Gibbs' Phase Rule provides the theoretical foundation, based in thermodynamics, for characterizing the chemical state of a (geologic) system.
  7. [7]
    [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.
  8. [8]
    [PDF] Thermodynamics of mixing
    This energy equation is a mean-field description of all binary regular ... is a miscibility gap between the two branches of the binodal in Fig. 4.8 ...
  9. [9]
    Entropy of Mixing - Chemistry LibreTexts
    Apr 1, 2023 · Δ ⁢ S = n ⁢ R ⁢ ln ⁡ V 2 V 1 . Now, for each gas, the volume ... Δ m ⁢ i ⁢ x ⁢ S = − R ⁢ ( n A ⁢ ln ⁡ χ A + n B ⁢ ln ...
  10. [10]
    miscibility gap (MT07270) - IUPAC
    A miscibility gap is observed at temperatures below an upper critical solution temperature (UCST) or above the lower critical solution temperature (LCST).Missing: consolute | Show results with:consolute
  11. [11]
  12. [12]
    13.2: Phase Diagrams- Binary Systems
    ### Summary of Binary Phase Diagrams with Miscibility Gaps
  13. [13]
    Thermoresponsive polymers in non-aqueous solutions
    Nov 10, 2022 · Fig. 1 Phase diagram showing the UCST and LCST for a binary mixture of a thermoresponsive polymer.
  14. [14]
    The lever rule
    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 ...Missing: miscibility gap UCST LCST solvus
  15. [15]
    [PDF] The Solubility of Triton X-114 and Tergitol 15-S-9 in High-Pressure ...
    Dec 8, 2005 · Changes in pressure can create a miscibility gap in an otherwise ... A miscibility gap exists for type-II binary systems at low temperatures (15.
  16. [16]
    Effect of pressure on liquidБ•'liquid miscibility gaps - AGU Journals
    In this way, a self-consistent thermodynamic database was developed for the binary phase diagram and its miscibility gap under any pressure. 6. Results and ...
  17. [17]
    [PDF] Calculation of ternary phase diagrams - eCommons
    The miscibility gap formed in a ternary solution is the result of Q 1-2? Q 1-3? or Q 2.3 being positive and large compared to. 2 RT, where T is some ...
  18. [18]
    Modelling of Ternary Alloy Systems Using Liquidus, Solidus, Solvus ...
    Aug 10, 2025 · The ternary alloy phase diagram is formed by an equilateral triangle, the vertices of which represent pure components.
  19. [19]
    Thermodynamic Modeling of Multicomponent Phase Equilibria
    The present paper gives an overview of the CALPHAD method and recent progress made. A brief history is given then the scope of phase diagram calculations is ...
  20. [20]
    An algorithm for calculating multicomponent miscibility gaps
    A general and very simple algorithm for calculating multicomponent miscibility gaps is described. Given one single expression for the free energy of a ...
  21. [21]
    Liquid immiscibility and core-shell morphology formation in ternary Al–Bi–Sn alloys
    ### Summary of Findings on Liquid Immiscibility in Ternary Al-Bi-Sn Alloys Leading to Core-Shell Microstructures
  22. [22]
    Core-shell microstructure formed in the ternary Fe-Co-Cu peritectic ...
    Apr 8, 2009 · The separated alloy melt rapidly solidifies and evolves core-shell microstructure composed of L 1 (Cu) and L 2 (Fe,Co) phases.
  23. [23]
    Nucleation transitions in undercooled Cu 70 Co 30 immiscible alloy
    Jul 31, 2014 · The present results indicate that homogeneous nucleation occurs and that the Cu-rich liquid phase has a retarding effect on the heterogeneous ...
  24. [24]
    [PDF] Theory of nucleation and growth during phase separation
    We present a new model for the entire process of phase-separation that combines steady-state homogeneous nucleation theory with the classical ...
  25. [25]
    [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).
  26. [26]
    Surface energy driven miscibility gap suppression during nucleation ...
    Jul 6, 2021 · The reason for the miscibility gap is a net repulsion between dissimilar atoms and is, in the regular solution model, described by the ...
  27. [27]
    Determination of the Metastable Zone Width and Nucleation ... - MDPI
    Sep 7, 2021 · The MSZW is, amongst others, affected by the rate of supersaturation generation, impurities and the surface available for heterogeneous ...
  28. [28]
    Atom Probe Study of the Miscibility Gap in CuNi Thin Films and ...
    While at the two higher temperatures total mixing is observed, a clear experimental evidence is found for a miscibility gap at 573 K with the boundary ...
  29. [29]
    [1611.07068] On the miscibility gap of Cu-Ni system - arXiv
    Nov 21, 2016 · The maximum of the symmetrical miscibility curve is around 800~K at Cu50%Ni50%. To our best knowledge, this is the first experiment determining ...
  30. [30]
    Microstructure of Alnico Alloys - AIP Publishing
    Due to a distinctly asymmetric form of the miscibility gap in the Fe‐NiAl system, the relative amounts of α and α′ are very temperature dependent.
  31. [31]
    Heat treatment of aluminum, Part V | Thermal Processing Magazine
    Feb 15, 2021 · Guinier-Preston (GP) zones are solute-rich clusters resulting from phase separation or precipitation within a metastable miscibility gap in the ...
  32. [32]
    Age-hardening in a Gold–Nickel Alloy | Nature
    In these systems complete miscibility occurs at high temperatures, but at lower temperatures there is a miscibility gap.
  33. [33]
    Leveraging miscibility gap via low-temperature annealing for ...
    Sep 17, 2025 · This study delves into the miscibility gap present in the phase diagram to tune nanoscale microstructure through controlled annealing for ...
  34. [34]
    Phase diagrams and microstructures related to the miscibility gap of ...
    Jul 20, 2025 · Two extensive liquid miscibility gap regions are observed in the ternary system. •. Core-shell and dispersion microstructures are found in ...
  35. [35]
    Age-hardening by miscibility limit in a multi-purpose dental gold ...
    This study examined the age-hardening by miscibility limit in a multi-purpose dental gold alloy containing platinum. The hardness increased rapidly in the.Missing: gap | Show results with:gap
  36. [36]
    Review Effect of 475 °C embrittlement on the mechanical properties ...
    May 20, 2009 · °C due to the presence of miscibility gap in iron–chromium binary alloy system embrittles the microstructure. Since this problem was ...
  37. [37]
    Low-temperature embrittlement in Fe–Cr–X (X = Ni, Al, Co, etc ...
    Aug 5, 2025 · The low-temperature embrittlement limits the service temperature of ferritic and duplex stainless steels. The effects of alloying elements ...
  38. [38]
    Crystallographic investigation of the huttenlocher exsolution at high ...
    Three plagioclases in the composition range An66–70 from the different geological environments were selected for investigation of the Huttenlocher ...
  39. [39]
    Revisiting the Bøggild Intergrowth in Iridescent Labradorite Feldspars
    Jul 5, 2021 · The Bøggild intergrowth occurs in the middle of the plagioclase solid solution, where the subsolidus phase relation is most complicated. Four ...
  40. [40]
    The peristerite gap in low-grade metamorphic rocks
    The variation of plagioclase composition in progressive metamorphic zones is explained by intersection of a plagioclase-forming reaction and the peristerite ...
  41. [41]
    On the Curie temperature and Nishizawa horn in the Al-Co-Ni system
    Mar 30, 2019 · Nishizawa et al. Thermodynamic analysis of solubility and miscibility gap in ferromagnetic alpha iron alloys. Acta Metall. (1979). F. Stein et ...Missing: Nishikawa | Show results with:Nishikawa
  42. [42]
    Molecular dynamics simulation of Al/Si-ordered plagioclase feldspar
    Mar 9, 2017 · In addition, because the peristerite, Bøggild, and Huttenlocher gaps did not appear in the phase relations determined from our simulations, it ...
  43. [43]
    Oiling Out in Crystallization - Mettler Toledo
    A LLPS due to a miscibility gap is a thermodynamic phenomenon (depending on system temperature, concentration and miscibility capability) and phase separation ...
  44. [44]
    Liquid–liquid phase separation in pharmaceutical crystallization
    Liquid–liquid phase separation (LLPS), or 'oiling out', is being reported increasingly during crystallization development of active pharmaceutical ingredients ( ...
  45. [45]
    Oiling Out - LinkedIn
    Nov 19, 2018 · This is called liquid-liquid phase separation, or oiling out, and is due to a miscibility gap in the phase diagram of the solute-solvent system.Missing: mixtures | Show results with:mixtures
  46. [46]
    Adjusting the low critical solution temperature of poly(N-isopropyl ...
    Major effects were also observed by changing the polarity of the bulk: the presence of 10% ethanol shifted the LCST of PNIPAM solution to lower temperatures, ...<|separator|>
  47. [47]
    UCST transitions in ethanol–water mixtures - RSC Publishing
    In aqueous solution copolymers exhibited a lower critical solution temperature (LCST) which decreased linearly with an increasing content of phenyl ethers.
  48. [48]
    Blend Miscibility and the Flory−Huggins Interaction Parameter
    Aug 6, 2025 · The Flory-Huggins interaction parameter quantitatively measures the miscibility of polymer blends [102, 103]. The following equation calculates ...
  49. [49]
    Predicting the Flory-Huggins χ Parameter for Polymers with Stiffness ...
    Jun 22, 2016 · The Flory–Huggins χ parameter describes the excess free energy of mixing and governs phase behavior for polymer blends and block copolymers.Missing: miscibility | Show results with:miscibility
  50. [50]
    Nucleation and spinodal decomposition of liquid mixtures
    Mar 1, 2005 · In this paper, we study the transition from states of unstable to states of metastable equilibrium of a liquid binary mixture that is quenched ...
  51. [51]
    Spinodal Decomposition - an overview | ScienceDirect Topics
    Spinodal decomposition is defined as a phase transition process that occurs within a miscibility gap, characterized by an unstable system against small ...
  52. [52]
    miscibility and potential applications of green solvents
    Green solvents play a central role in the design of more sustainable chemical processes. Among the properties influencing solvent selection, miscibility is ...
  53. [53]
    (PDF) The Solvent Miscibility Table Updated - ResearchGate
    Oct 7, 2025 · Green solvents play a central role in the design of more sustainable chemical processes. Among properties influencing solvent selection, ...
  54. [54]
    Rayleigh-Taylor instability in binary fluids with miscibility gap
    Aug 21, 2025 · Binary fluids with miscibility gap exhibit property variation in response to a stimuli, making them useful for applications like targeted ...
  55. [55]
    Rayleigh-Taylor instability in binary fluids with miscibility gap - arXiv
    Nov 25, 2024 · A novel phase field method is proposed to model the continuous transition of binary fluids exhibiting temperature sensitive miscibility gap.Missing: studies | Show results with:studies
  56. [56]
    Kelvin-Helmholtz instability in binary fluids with miscibility gap
    Sep 7, 2025 · In this study, a modified phase-field approach that allows modeling of partially miscible fluids and continuous transition from an initially ...
  57. [57]
    (PDF) Miscibility Gap Alloys: A New Thermal Energy Storage Solution
    Miscibility gap alloys (MGAs) are a new thermal storage technology that utilises the latent heat of fusion of metals. MGAs encapsulate the liquid phase while ...
  58. [58]
    Miscibility gap alloys with inverse microstructures and high thermal ...
    New high energy-density thermal storage materials are proposed which use miscibility gap binary alloy systems to operate through the latent heat of fusion ...
  59. [59]
    [PDF] Thermal capacitors made from Miscibility Gap Alloys (MGAs)
    Considering the mass fractions of the fusible phases the latent heat storages are 30.9 kJ (AlSn) and 187.6 kJ (FeCu) per kilogram of MGA. Accordingly, the iron- ...
  60. [60]
    Thermal Storage for CSP with Miscibility Gap Alloys - AIP Publishing
    The miscibility gap alloys are thermal storage materials which address many of the current thermal storage issues primarily through the use of metallic ...
  61. [61]
    Conceptual design and numerical analysis of a miscibility gap alloy ...
    May 10, 2024 · Miscibility gap alloy is a new type of shape-stabilized composite phase change material, which has the advantages of high energy storage density ...
  62. [62]
    Green steam and thermal storage startup reboots pilot plant after ...
    Apr 29, 2025 · Storage startup MGA Thermal is back on track and looking for capital to expand, after a hiccup in 2023 knocked it off course.
  63. [63]
    Energy Storage Demonstration Unit - MGA Thermal
    2025. April – We're thrilled to announce the successful commissioning of our demonstration plant - enabling continuous, cost-effective renewable steam ...Missing: RenewEconomy | Show results with:RenewEconomy
  64. [64]
    Further developments of the miscibility gap alloy: applications and ...
    May 9, 2025 · The practicality of implementing latent heat thermal storage has been problematic, mainly due to lower heat transfer rates due to low thermal ...
  65. [65]
    (PDF) Price estimation for miscibility gap alloy thermal storage systems
    Aug 6, 2025 · Miscibility gap alloys (MGAs) are an emerging thermal energy storage material with unique thermal properties that may be of particular ...<|separator|>
  66. [66]
    [PDF] Extended Thermal Cycling of Miscibility Gap Alloy High Temperature
    Abstract. 10. In previous work, a metastable miscibility gap was identified and exploited in the C-Al binary phase. 11 diagram, with 50% by volume of either ...
  67. [67]
    Tailoring the room-temperature miscibility gap in ordered spinel LiNi ...
    Jun 25, 2025 · The results demonstrate that multi-element doping enhances structural stability, tailored charge/discharge reaction mechanisms, and improves ...
  68. [68]
    Tailoring the room-temperature miscibility gap in ordered spinel LiNi ...
    This study investigates the effect of multiple element doping (Si, Ti, and Ge) on charge/discharge reaction mechanisms of LNMO cathodes to mitigate the capacity ...Missing: suppress separation
  69. [69]
    Mind the miscibility gap: Cation mixing and current density driven ...
    Nov 13, 2022 · Cathode materials that exhibit phase transitions with large structural rearrangements during electrochemical cycling are generally seen as disadvantageous.
  70. [70]
    [PDF] Mind the miscibility gap: Cation mixing and current density driven ...
    Nov 14, 2022 · For all samples the miscibility gap between the intermediate phases narrowed with increased current density, suggesting that the solid solution ...Missing: 2022-2025 | Show results with:2022-2025
  71. [71]
    Functional Electrolytes: Game Changers for Smart Electrochemical ...
    Oct 24, 2021 · Temperature-responsive electrolyte with reversible phase separation for Li-ion batteries. a) Digital photographs showing the phase ...
  72. [72]
    Thermally responsive polymers for overcoming thermal runaway in ...
    Feb 17, 2023 · Several strategies based on thermoresponsive polymers have been explored to prevent or inhibit the thermal runaway of ESDs due to rapid, ...
  73. [73]
    Pushing the Boundaries of Multicomponent Alloy Nanostructures
    Jun 17, 2022 · Thermodynamic prediction of the immiscible boundary in a temperature–composition space (miscibility gap) plays a key role in designing precursor ...
  74. [74]
    Kelvin-Helmholtz instability in binary fluids with miscibility gap - arXiv
    Sep 2, 2025 · [Submitted on 2 Sep 2025]. Title:Kelvin-Helmholtz instability in binary fluids with miscibility gap. Authors:Anubhav Dubey, Sakir Amiroudine.
  75. [75]
  76. [76]
    Clarifying the Ti-V Phase Diagram Using First-Principles ... - arXiv
    Jun 21, 2025 · A leading hypothesis attributes the miscibility gap to oxygen contamination during alloy preparation. To resolve this disagreement, we use ...Missing: methods | Show results with:methods
  77. [77]
    miscibility and potential applications of green solvents
    Sep 2, 2025 · Solvent miscibility is an important criterion in the selection of solvents for a process. While miscibility tables have been developed to guide ...