Fact-checked by Grok 2 weeks ago

Phase-field model

The phase-field model is a diffuse-interface approach in computational that simulates the spatiotemporal evolution of microstructures, such as phase transformations, , and , by representing phase boundaries as continuous transition zones rather than explicit sharp interfaces. This method reformulates moving-boundary problems into partial differential equations on fixed computational domains, enabling the handling of complex topological changes without explicit interface tracking. Historically, phase-field modeling traces its origins to J.D. van der Waals' 1894 theory of capillarity, which introduced the concept of diffuse interfaces to describe fluid phase equilibria, and was later formalized through the Ginzburg-Landau theory of order parameters in the 1950s. Key developments in the 1950s and 1970s, including John W. Cahn's work on , the derivation of the Cahn-Hilliard equation by Cahn and John E. Hilliard in 1958, and the Allen-Cahn equation by Samuel M. Allen and Cahn in 1972, provided the thermodynamic foundation for non-equilibrium dynamics. A major breakthrough occurred in 1993 with Ryo Kobayashi's phase-field model for dendritic solidification, which demonstrated practical simulations of morphological instabilities. At its core, the phase-field method relies on a free energy functional that incorporates bulk, gradient, and interfacial contributions, minimized through relaxational dynamics to drive microstructure evolution. For non-conserved order parameters, such as phase fractions, the Allen-Cahn equation governs evolution: \partial \phi / \partial t = -L (\delta F / \delta \phi), where \phi is the phase-field variable, L is the mobility, and F is the total free energy. For conserved quantities like composition, the Cahn-Hilliard equation applies: \partial c / \partial t = \nabla \cdot (M \nabla (\delta F / \delta c)), with M as the interfacial mobility, ensuring mass conservation across diffuse interfaces of finite width \epsilon. Extensions like the multi-phase-field model accommodate multiple phases and incorporate long-range effects, such as elasticity or fluid flow. Phase-field models are widely applied to predict phenomena in materials processing, including solidification microstructures in alloys, solid-state phase transformations like formation in steels, and ferroelectric domain structures in ceramics. They have enabled discoveries such as lattices in ferroelectric superlattices and enhanced piezoelectric responses in lead-based crystals through simulated domain engineering. Beyond materials, the approach extends to multiphase flows, tumor growth, and crack propagation, offering advantages in scalability and integration with .

Introduction

Definition and Core Concepts

The phase-field model is a continuum-based computational approach for simulating interfacial phenomena and phase transformations in materials, where interfaces between distinct s are represented as diffuse regions of finite width rather than infinitesimally thin sharp boundaries. This diffuse-interface framework originates from classical theories of nonuniform systems, allowing the evolution of complex microstructures to be described through partial differential equations without the need for explicit geometric tracking of boundaries. At its core, the phase-field model employs an order parameter, denoted as \phi(\mathbf{x}, t), which serves as a continuous scalar field varying smoothly across the interface to distinguish between phases—for instance, \phi = 1 in one phase (e.g., solid), \phi = 0 in another (e.g., liquid), and intermediate values in the transition zone. The dynamics are governed by a free energy functional that encapsulates the system's thermodynamic state, typically comprising bulk free energy terms favoring phase separation and gradient energy terms penalizing sharp variations to enforce interface smoothness. Evolution of the order parameter proceeds via relaxational dynamics that minimize this free energy, often derived from variational principles rooted in nonequilibrium thermodynamics. Compared to sharp-interface methods, phase-field models offer significant advantages, including the automatic accommodation of topological changes such as interface merging or pinching without interventions, and their suitability for resolving intricate morphologies in multiphase systems. Physically, this approach is motivated by thermodynamic principles, particularly the minimization of in systems with interfacial contributions, extending early square-gradient theories to capture realistic interface energies and phase stability.

Historical Development

The foundations of phase-field modeling trace back to early theories of diffuse interfaces, with J.D. van der Waals introducing the concept in 1894 through his analysis of atomic forces and the structure of fluid interfaces, laying the groundwork for non-sharp boundary descriptions. In the 1950s and 1960s, this idea evolved into more rigorous thermodynamic frameworks, notably through the work of John W. Cahn and John E. Hilliard, who in 1958 derived an expression for the of nonuniform systems, incorporating gradient terms to model interfacial energy and enabling the description of in alloys via what became known as the Cahn-Hilliard equation. Their 1959 follow-up further refined these models for processes, emphasizing conserved order parameters like composition. By the 1970s, related developments included the Allen-Cahn equation (1979), which addressed non-conserved order parameters for phase ordering, providing a complementary tool for interface motion without mass conservation. The 1980s marked the pioneering application of phase-field approaches to free-boundary problems like solidification, with Anthony J. Fix proposing early models for diffuse interfaces in moving boundary contexts around 1980, though much of his work appeared in proceedings by 1983. Independently, J.S. Langer advanced phase-field formulations for in dendritic solidification, developing models in the mid-1980s that captured instabilities in without explicit interface tracking. Gunduz Caginalp contributed significantly by deriving phase-field equations in the 1980s and demonstrating their convergence to sharp-interface limits, such as the , through , thus bridging diffuse and classical models. In the , phase-field methods expanded rapidly for quantitative simulations of complex microstructures, with Alain Karma and Wouter-Jan Rappel introducing refined models for dendritic growth that incorporated thin-interface asymptotics to match sharp-interface predictions accurately. Variants of the Allen-Cahn equation were adapted during this period to handle anisotropic growth and solute trapping in alloys, enhancing the model's versatility for solidification dynamics. Multiphase extensions emerged, such as Steinbach et al.'s 1996 model, which used multiple order parameters to simulate motion and phase competitions. From the onward, phase-field modeling shifted toward computational implementation and broader applications, integrating multiphase frameworks for polycrystalline materials and extending to fields like for dynamics. A seminal review by Boettinger et al. in 2002 synthesized these advances, highlighting phase-field's role in microstructure evolution across solidification and solid-state transformations. This era also saw a milestone transition from analytical to numerical focus, with finite element methods gaining prominence around 2004 for handling adaptive meshes and complex geometries in large-scale simulations.

Mathematical Formulation

Basic Phase-Field Equations

The phase-field model describes the evolution of an order parameter φ, which varies continuously across interfaces between distinct phases, typically taking values such as φ = -1 in one phase and φ = +1 in the other. The foundational dynamics are governed by the functional F[\phi] = \int_V \left[ f(\phi) + \frac{\varepsilon^2}{2} |\nabla \phi|^2 \right] dV, where f(φ) represents the bulk density, and the gradient energy term with interface width parameter ε penalizes sharp variations in φ, effectively smearing the interface over a diffuse region of thickness ~ε. This functional, rooted in Ginzburg-Landau theory, provides the thermodynamic driving force for phase evolution. For non-conserved order parameters, such as in solid-solid phase transformations where volume fractions are fixed, the evolution follows the Allen-Cahn equation: \frac{\partial \phi}{\partial t} = -M \frac{\delta F}{\delta \phi}, with mobility M determining the rate of motion. Here, the δF/δφ = -ε² ∇²φ + df/dφ acts as a generalized , driving φ toward minima of F and resulting in motion to itself at a velocity proportional to local , which models dissipative coarsening and ordering processes. In contrast, for conserved order parameters, as in binary alloy where atomic diffusion maintains , the Cahn-Hilliard ensures : \frac{\partial \phi}{\partial t} = \nabla \cdot \left( [M](/page/M) \nabla \frac{\delta F}{\delta \phi} \right), where the μ = δF/δφ, and the is J = - ∇μ. This fourth-order PDE captures diffusion-driven and coarsening, with interfaces migrating to reduce bulk differences while preserving ∫φ dV. These equations often couple with additional fields; for instance, in solidification, φ evolves alongside temperature T via ∂T/∂t = (latent heat)/c_p * ∂φ/∂t + diffusion terms, where c_p is specific heat, linking thermal transport to phase change. Similarly, concentration fields can couple for multicomponent systems, extending the model to realistic material processes. Typical simulations employ to mimic bulk behavior without surface effects, or no-flux conditions (∇φ · n = 0, ∇μ · n = 0) for insulated domains, with initial conditions setting φ as step functions smoothed over ε to seed . These setups facilitate numerical resolution of the diffuse interface while capturing long-time microstructural dynamics.

Variational Principles

The phase-field model is fundamentally derived from thermodynamic variational principles, where the evolution of the order parameter \phi is formulated as a gradient flow that minimizes a functional F[\phi]. This functional typically comprises bulk and interfacial contributions, such as F[\phi] = \int_\Omega \left( f(\phi) + \frac{\epsilon^2}{2} |\nabla \phi|^2 \right) dx, ensuring thermodynamic consistency. The dynamics follow from the principle of least action combined with energy dissipation, leading to evolution equations that decrease F over time. In the Allen-Cahn formulation, suitable for non-conserved order parameters like spinodal decomposition, the evolution is a L^2-gradient flow: \partial_t \phi = -M \frac{\delta F}{\delta \phi}, where M > 0 is the mobility and \frac{\delta F}{\delta \phi} = -\epsilon^2 \Delta \phi + f'(\phi) is the functional derivative. For the Cahn-Hilliard equation, which conserves the order parameter (e.g., in binary alloy phase separation), the flow is in the H^{-1} metric to enforce mass conservation: \partial_t \phi = M \Delta \frac{\delta F}{\delta \phi}, resulting in a fourth-order PDE. These derivations stem from the Rayleigh dissipation functional, which quantifies irreversible energy loss as \mathcal{D}[\dot{\phi}] = \frac{1}{2} \int_\Omega \frac{1}{M} (\partial_t \phi)^2 \, dx for Allen-Cahn or analogous forms involving fluxes for Cahn-Hilliard, minimized alongside the rate of change of F. The Onsager reciprocal relations further underpin this by linking dissipative fluxes to thermodynamic forces via symmetric mobility matrices, ensuring the framework applies to non-equilibrium processes. The derivation proceeds by postulating the time evolution as the negative of F in an appropriate , with M scaling the rate and reflecting microscopic coefficients. For instance, setting the variational derivative to drive the yields the law directly, while the choice of ( L^2 vs. H^{-1} ) determines whether the are relaxational or diffusive. In coupled systems, such as those incorporating fluid advection or deformations, the extends by including cross-terms in the (e.g., \phi to or fields) and modifying the functional to for interactions, like advective fluxes \mathbf{J} = -M \nabla \mu + \phi \mathbf{u}. This maintains a unified structure across multiphysics problems. Theoretically, these principles justify the model's consistency with the second law of thermodynamics, as the energy dissipation rate satisfies \frac{dF}{dt} = -\int_\Omega \frac{1}{M} \left( \partial_t \phi \right)^2 \, dx \leq 0 for Allen-Cahn, with analogous non-positive forms for Cahn-Hilliard, ensuring monotonic decrease and thermodynamic irreversibility.

Energy Density Functions and Alternatives

The functional in phase-field models incorporates a bulk density that drives and a term that accounts for the cost of interfaces. A standard choice for the bulk in systems is the f(\phi) = \frac{1}{4} \phi^2 (1 - \phi)^2, where the phase-field variable \phi interpolates between 0 and 1, with minima at these values representing the two phases. This quartic form ensures equal depths at the wells, promoting symmetric phase coexistence. The accompanying gradient energy density is typically \frac{\varepsilon^2}{2} |\nabla \phi|^2, which introduces a length scale for diffuse interfaces and contributes to the excess free energy at boundaries between phases. The resulting interfacial energy \sigma for a planar interface is obtained by integrating across the transition profile: \sigma = \int_{0}^{1} \varepsilon \sqrt{2 f(\phi)} \, d\phi, yielding a value proportional to \varepsilon for the given potential. Alternatives to the polynomial double-well potential include sinusoidal forms, which are employed in phase-field crystal models to capture periodic atomic density variations rather than simple binary separation. For instance, the potential may incorporate terms like -\cos(2\pi \phi) to model oscillatory free energies in crystalline structures. In crystal growth simulations, anisotropic energy densities modify the gradient term to reflect orientation-dependent interfacial properties, often as \frac{\varepsilon^2}{2} a(\mathbf{n}) |\nabla \phi|^2, where a(\mathbf{n}) is a function of the interface normal \mathbf{n} derived from symmetry considerations. Non-local variants extend the energy density to include integral operators over spatial domains, enabling the incorporation of long-range elastic or diffusive interactions in more complex systems. The parameter \varepsilon acts as a regularizing length scale that controls the artificial thickness of the diffuse , typically selected on the order of several spacings to ensure while approximating physical sharpness. Larger \varepsilon values reduce computational cost by coarsening the interface but can alter kinetics, such as slowing solute trapping during rapid solidification, whereas smaller values enhance fidelity to sharp-interface dynamics at the expense of finer meshes. At , the one-dimensional profile across an minimizes the , yielding the hyperbolic tangent solution \phi(z) = \frac{1}{2} \left( 1 + \tanh\left( \frac{z}{\varepsilon \sqrt{2}} \right) \right) for the standard , where the transition width is proportional to \varepsilon. This profile ensures that the interfacial energy matches the integrated excess , linking the model parameter directly to observable widths in simulations.

Theoretical Extensions

Sharp Interface Limit

The sharp interface limit of phase-field models occurs as the interface thickness parameter ε approaches zero, wherein the diffuse transition region between phases shrinks to a mathematical surface, recovering classical descriptions of phase boundaries governed by free-boundary problems. Asymptotic analyses, such as matched asymptotic expansions and front-tracking methods, reveal that the evolution of the phase-field variable φ leads to an interface motion law of the form V = -\kappa, where V is the normal velocity of the and \kappa is its , corresponding to for the pure Allen-Cahn equation. In coupled systems, this limit incorporates additional effects like the Gibbs-Thomson relation, which modifies the equilibrium as T_i = T_m - \Gamma \kappa, with \Gamma denoting the interfacial energy and T_m the , alongside solvability conditions for the underlying fields. For applications in solidification, early asymptotic studies demonstrated convergence to the , where the interface velocity satisfies V = -\frac{1}{L} \mathbf{n} \cdot (D_s \nabla T_s - D_l \nabla T_l), with L the , D_{s,l} the diffusivities in and phases, and \mathbf{n} the normal vector, coupled to diffusion equations away from the . These analyses highlighted discrepancies arising from finite ε, such as spurious contributions to interfacial and . To mitigate this, thin-interface asymptotics were developed in the , enabling corrections for unequal conductivities and allowing ε to be comparable to the capillary length d_0 = \Gamma / L without severe loss of accuracy, thus facilitating efficient numerical simulations of dendritic growth. Rigorous convergence proofs establish that, under conditions such as sufficiently small ε and , solutions to the phase-field equations converge in appropriate norms to weak solutions of the sharp-interface Stefan or Mullins-Sekerka problems, which describe unstable morphological evolution driven by diffusion instabilities. These proofs typically employ energy estimates and compactness arguments to pass to the limit, confirming the recovery of curvature-driven interface conditions and across the free boundary. Despite these theoretical foundations, the sharp interface limit imposes limitations, particularly in modeling dendritic growth, where phase-field introduce artificial kinetic coefficients that alter the interface mobility and selection of stable patterns beyond the leading-order asymptotics. Additionally, solvability conditions—required in sharp-interface theories to resolve the non-uniqueness of dendritic solutions—manifest implicitly in phase-field models through parameter choices, but finite ε can lead to deviations that necessitate higher-order for quantitative .

Multiphase-Field Models

Multiphase-field models extend the phase-field approach to systems involving more than two phases or components by introducing multiple order parameters \phi_i, where i = 1, 2, \dots, N labels the distinct phases, satisfying the \sum_i \phi_i = 1. This allows for the diffuse of multiple interfaces and their junctions, simulations of microstructural without explicit interface tracking. The approach originated in the mid-1990s as a means to handle multi-phase transformations in alloys, building on the single-order-parameter framework by generalizing the functional to accommodate pairwise interactions between phases. The total free energy F in multiphase-field models is typically expressed as an integral over the domain: F = \int \left[ \sum_i \phi_i f_i(c) + \sum_i \frac{\epsilon^2}{2} |\nabla \phi_i|^2 + \sum_{i<j} W_{ij}(\phi_i, \phi_j) \right] dV, where f_i(c) are the bulk free energies of each phase (potentials local to each phase depending on composition c), \epsilon controls the interface width, and W_{ij}(\phi_i, \phi_j) are double-well or barrier potentials enforcing phase separation and calibrated such that the interface energy between phases i and j is \sigma_{ij}. This structure ensures that the model recovers sharp-interface limits for binary interactions while handling triple or higher junctions naturally through the collective order parameters. The gradient terms promote smooth transitions at interfaces, and anisotropy can be incorporated by making \epsilon or W_{ij} orientation-dependent. Interface energies \sigma_{ij} are defined pairwise, allowing anisotropic or orientation-dependent properties by incorporating terms like \sigma_{ij}(\mathbf{n}_{ij}), where \mathbf{n}_{ij} is the interface normal. To avoid negative or non-physical order parameter values that could lead to spurious phases, barrier functions such as W_{ij} = \lambda_{ij} g(\phi_i, \phi_j) are included, with g a smooth double-well function (e.g., g = \phi_i^2 \phi_j^2) and \lambda_{ij} a height parameter tuned to match \sigma_{ij}. A global barrier like W = \lambda g(\{\phi_i\}) \left(1 - \sum_i \phi_i \right)^2 with g = \sum_i \phi_i^2 (1 - \phi_i)^2 is also used to enforce the constraint. These barriers maintain the convexity of the free energy landscape and ensure thermodynamic consistency at multi-phase junctions. The temporal evolution of the order parameters follows coupled : \tau_i \frac{\partial \phi_i}{\partial t} = -M_i \frac{\delta F}{\delta \phi_i} + \sum_{j \neq i} \Lambda_{ij} \left( \frac{\delta F}{\delta \phi_i} - \frac{\delta F}{\delta \phi_j} \right), where \tau_i and M_i are relaxation times and mobilities for phase i, and \Lambda_{ij} are Lagrange multipliers enforcing \sum_i \phi_i = 1. This coupling drives the minimization of F, with interfaces migrating normal to themselves at velocities proportional to the local curvature and energy gradients. For diffusive processes, solute conservation is coupled via an advection-diffusion equation with anti-trapping corrections to mitigate artificial solute piling at interfaces. These models find applications in simulating grain growth in polycrystals, where each grain is a phase and \sigma_{ij} reflects grain boundary energies, leading to realistic triple-junction dynamics as validated in benchmark studies. In ternary alloys, the framework captures eutectic solidification patterns by treating liquid and two solid phases separately, predicting lamellar or rod-like microstructures under controlled undercooling. Key challenges include enforcing the normalization constraint efficiently, often via projection methods or iterative solvers for the Lagrange terms, which can introduce numerical stiffness. Computational cost scales quadratically with the number of phases (O(N^2)) due to pairwise couplings, limiting large-scale simulations without adaptive meshing or parallelization; for instance, systems with over 100 phases require specialized implementations to remain tractable.

Adaptations to Non-Standard Domains

Phase-field models, traditionally formulated on continuous Euclidean domains, have been adapted to non-standard, discrete structures such as graphs and lattices to address scenarios where spatial continuity does not apply, including networked systems and irregular geometries. These adaptations redefine the core components—such as the order parameter evolution and free energy functionals—in terms of discrete operators, enabling simulations on combinatorial structures like node-edge graphs or grid-based lattices. In graph-based phase-field models, nodes represent sites analogous to spatial points, while edges encode interactions that approximate gradients and Laplacians. The evolution of the phase-field variable \phi at each node follows a discrete Allen-Cahn equation driven by a graph Laplacian, which promotes smoothing and phase separation across the network. This framework has been applied to model dynamics on networks, such as flow propagation or collective behaviors in social systems, where the graph topology dictates the interface motion. A key adaptation involves reformulating the free energy functional on the graph. The interfacial energy term, which penalizes differences in \phi across phase boundaries, is expressed as a weighted sum over edges: E_{\text{interface}} = \frac{\epsilon}{2} \sum_{(i,j) \in \mathcal{E}} w_{ij} (\phi_i - \phi_j)^2, where \epsilon > 0 is a small parameter controlling interface width, w_{ij} are edge weights reflecting connectivity strength, and \mathcal{E} denotes the edge set. Combined with a bulk potential term summed over nodes, this discrete energy drives the phase-field dynamics via gradient flow, yielding a variational structure akin to the continuum case. Lattice models integrate phase-field principles with discrete frameworks like cellular automata or finite-difference schemes on meshes, particularly useful for handling irregular boundaries and heterogeneous structures. In these adaptations, the phase-field order parameter evolves on lattice sites using local update rules that incorporate diffusion-like operators, such as discrete Laplacians defined via nearest-neighbor interactions. This approach facilitates boundary treatments in complex domains, for instance, by incorporating site-specific potentials to model exclusions or barriers, as seen in simulations of flow through porous structures. Convergence to the continuum limit is established for these discrete models under suitable conditions, such as densities approximating manifolds or spacings approaching zero. For -based variants, as the refines (e.g., via increasing density and edge weights mimicking metric distances), the discrete and phase evolution recover and Allen-Cahn dynamics in the . Similar asymptotic analyses hold for discretizations, ensuring consistency with underlying partial equations. Representative examples include modeling percolation transitions on graphs, where phase-field evolution captures the emergence of connected clusters as a critical phase change, with the order parameter indicating site occupancy. Likewise, epidemic spreading on networks can be framed as a phase transition, with the graph phase-field simulating the propagation of infection states across edges, analogous to interface growth in continuous media. These discrete adaptations extend naturally to multiphase settings by introducing multiple order parameters per node.

Applications in Materials Science

Solidification and Phase Transformations

Phase-field models have been extensively applied to simulate solidification processes, where a liquid transforms into a phase, often leading to complex microstructural patterns such as dendrites. These models integrate the evolution of a , φ, which smoothly transitions from 0 in the to 1 in the , with governing equations for and solute concentration to capture the coupled of and mass transport during phase change. In pure materials, the model couples the phase-field equation to the diffusion equation, ∂T/∂t = D_T ∇²T - (L / c_p) ∂φ/∂t, where T is , D_T is , L is the , and c_p is , allowing the interface to advance based on undercooling. For binary alloys, an additional solute conservation equation, ∂c/∂t = D_c ∇²c + advection terms if flow is present, is incorporated, with c representing concentration; the phase-field incorporates thermodynamic driving forces from both thermal and solutal undercooling to model solute rejection at the interface. In dendritic growth during solidification, phase-field models naturally capture the morphological instability of planar interfaces, known as the Mullins-Sekerka instability, where perturbations amplify due to diffusion fields, leading to branched structures. in the interfacial energy, introduced via orientation-dependent terms in the phase-field functional, such as ε(∇φ)^2 with ε varying with the interface normal, selects stable dendritic patterns and determines velocity and radius through solvability conditions derived from marginal stability theory. These models, refined through thin-interface asymptotics to eliminate unphysical effects like solute trapping in the diffuse , accurately predict growth matching the sharp-interface limit for low undercooling. For eutectic and peritectic transformations in , multiphase-field extensions allow simulation of coupled growth of multiple solid phases from the liquid, critical for processes where dictates phase selection. In eutectic solidification, the model resolves lamellar or rod-like patterns emerging from cooperative growth, with interface energies and diffusivities governing spacing and stability; for peritectic reactions, it captures the transient liquid-to-solid transformation, including peritectic phase formation ahead of primary dendrites. These simulations highlight how constitutional undercooling drives , aiding optimization of microstructures in industrial . Validation of phase-field models against experiments has been achieved through comparisons with in transparent alloys like succinonitrile-acetone, where measured dendritic spacings and velocities from 1980s microgravity and ground-based setups align with simulations within 10-20% for undercoolings up to 1 K. Similar benchmarks using or TRIS-NPG for peritectic systems confirm the models' ability to reproduce transition velocities and phase fractions observed in real-time optical experiments from the to . These agreements underscore the quantitative reliability of phase-field approaches when calibrated against the sharp-interface limit for interface .

Fracture Mechanics

In phase-field modeling of , cracks are represented as a diffuse transition zone via a phase-field φ, where φ = 1 denotes intact and φ = 0 indicates a fully cracked state. This approach regularizes sharp crack discontinuities into a continuous damage field, enabling the simulation of complex crack topologies without explicit tracking of interfaces. The foundational framework stems from the variational formulation of brittle , which minimizes a total functional comprising elastic and . A canonical form of the fracture energy functional is given by F = \int_{\Omega} \left[ \phi^2 \psi_e^+(\mathbf{u}) + G_c \left( \frac{(1 - \phi)^2}{2l} + \frac{l}{2} |\nabla \phi|^2 \right) \right] dV, where \psi_e^+(\mathbf{u}) is the positive (tensile) part of the elastic strain energy density dependent on the displacement field \mathbf{u}, G_c is the critical fracture energy (surface energy release rate), and l is the length scale parameter controlling the width of the diffuse crack. This functional approximates the Griffith energy for sharp cracks in the limit as l → 0, recovering the classical criterion for crack propagation when the energy release rate equals G_c. The evolution of the phase field is governed by an energy minimization principle adapted for fracture, where the decrease in φ (representing crack advance) is driven by the release of stored strain energy. The governing equation for φ arises from the Euler-Lagrange equation of the functional, typically resulting in a fourth-order : G_c \left( \frac{1 - \phi}{l} + l \Delta \phi \right) = 2 \phi \psi_e^+(\mathbf{u}), subject to boundary conditions and an irreversibility constraint φ ≤ φ_{n-1} (from the previous time step) to prevent crack healing, enforced via a . This constraint ensures monotonic damage progression, aligning with physical observations in brittle materials. In Miehe et al.'s formulation, the model incorporates a degradation function on the tensile part of the strain energy to avoid spurious crack growth under compression, enhancing robustness for rate-independent propagation. For ductile , the phase-field approach is extended by the with a model, where plastic dissipation contributes to the crack driving force alongside . This integration allows simulation of void growth, coalescence, and shear-dominated failure, with the phase field degrading both and contributions to the . The sharp-interface limit of these models recovers the Griffith modified for mixed-mode loading, enabling predictions of in metals. In contrast to purely brittle models, the ductile variants incorporate yield criteria like J2 , capturing post-yield softening and localization. Applications of phase-field fracture models extend to dynamic regimes, where inertial effects are included via the kinetic energy term in the , allowing simulation of high-speed crack branching and fragmentation in brittle materials like or ceramics. For , cyclic loading is modeled by accumulating damage over load cycles through history variables in the driving force, predicting crack initiation and growth under conditions in alloys. These formulations, notably advanced by Miehe and collaborators in the , have been validated against experimental benchmarks, demonstrating accurate recovery of paths in quasi-static and dynamic tests.

Microstructure Evolution

Phase-field models simulate the evolution of microstructure in polycrystalline materials by representing grains as diffuse interfaces within a multi-phase , where each phase field variable corresponds to the orientation of a grain. This approach captures the collective dynamics of grain boundaries, enabling predictions of changes, size distributions, and development without explicit tracking of sharp interfaces. The driving force for evolution typically arises from reductions in interfacial , leading to curvature-driven boundary migration. In multi-phase-field models for polycrystalline systems, grain boundary is incorporated as an orientation-dependent quantity, denoted as \sigma(\theta), where \theta represents the misorientation or inclination angle between adjacent grains. This influences and , with higher- boundaries migrating faster to minimize total . Curvature-driven motion is modeled through the evolution for phase fields, where the velocity of a is proportional to its local curvature \kappa and \sigma(\theta), resulting in the shrinkage of high-curvature grains and growth of larger ones. Such formulations accurately reproduce experimental observations of faceted or tilted boundaries in metals. Coarsening in phase-field simulations follows classical laws adapted to diffuse-interface descriptions, such as the Lifshitz-Slyozov-Wagner (LSW) theory for precipitate coarsening or Hillert's distribution for , where the average grain size R scales with time t as R \propto t^{1/2} or t^{1/3} depending on the controlling mechanism. These models validate the self-similar evolution of size distributions, with phase-field results converging to Hillert's skewed log-normal form for ideal in three dimensions after accounting for finite simulation sizes. The inclusion of impurities, modeled as dispersed second-phase particles, modifies coarsening by exerting Zener pinning forces that slow boundary motion and stabilize finer microstructures, as particles with volume fractions around 1-5% can reduce growth rates by factors of 2-10 compared to pure systems. Recrystallization processes are simulated by coupling phase-field variables with a coarse-grained field \rho, where stored deformation drives the and of strain-free grains. The evolution incorporates a driving term proportional to the dislocation difference across boundaries, \Delta E \approx \frac{1}{2} \mu b^2 \rho, with \mu as the and b the , leading to boundary velocities that depend on local \rho gradients. This framework distinguishes normal , where recrystallized grains expand uniformly to refine the matrix, from abnormal , where select large grains dominate due to localized high-mobility boundaries or reduced pinning, often resulting in bimodal size distributions observed in deformed metals. Seminal implementations demonstrate that abnormal initiates when dislocation densities exceed critical thresholds around $10^{14}-10^{15} m^{-2}, promoting rapid texture changes. Applications of these models include steel processing, where phase-field simulations predict grain coarsening during annealing, aiding optimization of heat treatments to achieve desired strengths in low-carbon steels by controlling final grain sizes to 10-50 \mum. In thin-film deposition, such as physical vapor deposition of metallic coatings, the models capture orientation-dependent growth and texture evolution on substrates, revealing how initial nucleation densities influence columnar structures and reduce defects in films 100-500 nm thick. Early work by Fan and Chen in the 1990s established two-dimensional phase-field frameworks for these processes, demonstrating realistic topological transitions in grain networks during isothermal holding.

Applications in Biological and Other Systems

Collective Cell Migration

Phase-field models have been adapted to simulate collective cell migration, where groups of cells coordinate their movement through processes such as , haptotaxis, and mechanical interactions, capturing the emergent behaviors observed in biological tissues. These models represent cell density using a phase-field \phi, which evolves diffusely to describe interfaces between cellular and non-cellular regions, enabling the simulation of complex topologies without explicit tracking of boundaries. A core feature in these models is the coupling of the cell density \phi with hydrodynamic velocity fields \mathbf{u}, often achieved by combining a Cahn-Hilliard equation for the conservative evolution of \phi with the Navier-Stokes equations for incompressible fluid flow. The Cahn-Hilliard component governs and of cell density, while by \mathbf{u} incorporates collective motion; directed is introduced via flux terms for (response to chemical gradients) or haptotaxis (response to gradients). This framework allows simulation of cells navigating heterogeneous environments, such as mazes or fibrous matrices, where collective speeds can increase due to enhanced propulsion from adhesions. Adhesion and proliferation are incorporated through a free-energy functional that includes interaction terms promoting cell-cell and cell-substrate , alongside double-well potentials to enforce volume constraints and for clustering. Repulsive and attractive potentials in the drive aggregation, mimicking cadherin-mediated adhesions, while proliferation is modeled by source terms dependent on \phi, leading to growth in high-density regions. mechanisms facilitate the formation of multicellular clusters, with aggregation thresholds observed at characteristic intercellular separations, highlighting how energetic penalties stabilize collective structures. In applications, phase-field models describe the motion of epithelial sheet interfaces, where the diffuse interface advances to close gaps, driven by active stresses and polarity alignment. Frameworks developed by Camley et al. in the integrate velocity-dependent free energies to capture leader-follower and supracellular alignment, reproducing experimental closure rates in 2D monolayers at speeds comparable to observed values. These models emphasize how interfacial and inhibition guide the propagation of fronts. Recent extensions as of model mechanosensitive migration on the surface and interior of morphing soft , enhancing understanding of . Multiphase-field extensions represent distinct types with multiple order parameters \phi_i, enabling simulations of and patterns through differential mobilities and interaction energies. For instance, varying strengths between phases can lead to of motile and non-motile cells, or invasive fronts where aggressive types penetrate quiescent tissues, as seen in tumor models with invasion depths scaling with chemotactic sensitivities up to 1000 nM/s rates. This approach reveals how differential and drive phase , with applications to embryonic and . As of 2025, multi-phase-field models have been advanced to handle deformable interfaces in biological tissues more efficiently.

Electrochemical Processes

Phase-field models have been extended to electrochemical processes to simulate the evolution of interfaces during , operation, and , where diffuse interfaces naturally capture morphological instabilities without explicit tracking. These models couple the phase-field order parameter φ, which distinguishes between phases such as and , with electrochemical transport equations to describe , , and reaction . A key approach integrates the phase-field formalism with the equations for transport and the Butler-Volmer kinetics for interfacial reactions, enabling quantitative predictions of deposition and dissolution rates. In , the Butler-Volmer equation governs the at the as a function of , with the phase-field φ modulating the reaction rate across the diffuse . The framework accounts for electrostatic potential and species concentrations, allowing the model to capture ion transport limitations and the resulting evolution. This coupling has been shown to recover sharp-interface kinetics in the thin-interface limit, providing a thermodynamically consistent description of processes like metal . For instance, nonlinear phase-field variants incorporate higher-order corrections to accurately reproduce Butler-Volmer behavior even under large , essential for systems far from . Dendritic electrodeposition poses significant challenges in lithium metal batteries, where uneven lithium plating leads to capacity loss and safety risks. Phase-field models address this by coupling the evolution of φ with the electric potential ψ, incorporating electromigration and diffusion of lithium ions via PNP equations. These simulations reveal how overpotential gradients drive morphological instabilities, promoting dendrite initiation and growth from planar fronts. In lithium batteries, the models predict that higher current densities exacerbate dendrite formation, with branching patterns influenced by electrolyte properties and electrode surface conditions. Quantitative studies using thin-interface asymptotics have validated these models against experimental dendrite velocities and tip radii in symmetric lithium cells. As of 2023, convergent non-oscillatory algorithms have improved predictions of shape changes in lithium metal batteries. Corrosion pitting involves localized at defects, where phase-field models simulate the transition from general to by coupling φ with and multi-phase . In multi-phase setups, the order parameter tracks the evolving morphology, incorporating ion diffusion and reaction rates to predict depth and shape over time. For metallic materials like iron or magnesium alloys, these models incorporate pH-dependent and effects to capture activation-controlled initiation and diffusion-limited . Simulations demonstrate how ions accelerate pitting by stabilizing local acidity, leading to irregular fronts that evolve into complex structures. Examples of phase-field applications in electrochemical contexts include the prediction of secondary arm spacing during solidification under electrochemical gradients, where solute and coarsening are modeled to match experimental spacings in Fe-C and Al-Cu systems. These simulations highlight how cooling rates influence arm spacing, providing insights into microstructure control in electrodeposited . Additionally, models inspired by Newman's porous electrode theory from the 2000s have been adapted to phase-field frameworks for interfaces, incorporating volume-averaged with interfacial to simulate suppression strategies. Recent quantitative phase-field models as of enhance simulations of electrochemical systems in .

Fluid Dynamics Interfaces

In phase-field modeling of interfaces, the diffuse interface representation is coupled with the incompressible Navier-Stokes equations to simulate multiphase flows such as droplet dynamics and immiscible fluid interactions. The phase-field variable \phi evolves according to either the non-conservative Allen-Cahn , \partial_t \phi + \mathbf{u} \cdot \nabla \phi = -M \left( \frac{\delta F}{\delta \phi} \right), or the conservative Cahn-Hilliard , \partial_t \phi + \nabla \cdot (\mathbf{u} \phi) = \nabla \cdot (M \nabla \mu), where \mu = \frac{\delta F}{\delta \phi} is the and F is the Ginzburg-Landau functional. These are combined with the momentum \rho (\partial_t \mathbf{u} + \mathbf{u} \cdot \nabla \mathbf{u}) = -\nabla p + \nabla \cdot (2 \eta D + \mathbf{T}_K) + \mathbf{f}, where \mathbf{u} is the velocity, \rho and \eta are and , D is the strain rate tensor, and \mathbf{f} includes body forces. Surface tension emerges naturally from the Korteweg stress tensor \mathbf{T}_K = \lambda (\nabla \phi \otimes \nabla \phi - \frac{1}{2} |\nabla \phi|^2 \mathbf{I} - \frac{1}{2} \phi \nabla^2 \phi \mathbf{I}), with \lambda related to interfacial tension \sigma, enabling the model to capture effects without explicit interface tracking. The conservative forms, particularly the Cahn-Hilliard equation, ensure mass conservation across phases, which is crucial for long-time simulations of volume-preserving flows, while non-conservative Allen-Cahn variants may require modifications like Lagrange multipliers to mitigate mass loss. In droplet dynamics, this facilitates the study of coalescence and in emulsions, where changes occur automatically as the diffuse thins and merges or fragments under hydrodynamic stresses. For instance, during coalescence, the phase-field allows formation between droplets without remeshing, contrasting with sharp-interface methods, and simulations reveal that conservative formulations better preserve droplet volumes during in shear flows compared to non-conservative ones, reducing artificial by up to 1% over extended times. For turbulent multiphase flows, phase-field models are integrated with large eddy simulations (LES) to resolve subgrid-scale effects while capturing interfacial instabilities. The filtered Navier-Stokes equations incorporate the phase-field for interface advection, with subgrid models for the Korteweg stress to handle unresolved capillary waves. This approach has been applied to Rayleigh-Taylor instability, where denser fluid penetrates lighter fluid under gravity, showing good agreement with experiments for Atwood numbers around 0.5, with growth rates matching linear stability theory within 5% error in early stages. Extensions draw from front-tracking methods, adapting phase-field for complex geometries in high-Reynolds-number regimes. Recent developments as of 2023 include conservative second-order phase-field models for n-phase simulations and improved lattice-Boltzmann integrations for high density ratios. Applications include , where phase-field simulations model droplet ejection and impact, capturing pinch-off and splashing dynamics driven by inertial and capillary forces, aiding optimization of printhead designs for resolutions below 10 \mum. In ocean wave modeling, adaptive phase-field methods simulate breaking waves, resolving and spray formation over kilometer-scale domains, with computational costs reduced by 50% via multiresolution grids compared to uniform meshes. These build on early multiphase simulation frameworks from the , extending them to diffuse interfaces for enhanced topology handling in real-world hydrodynamic problems.

Numerical Implementation

Discretization and Solution Methods

The phase-field model equations, typically comprising higher-order partial differential equations such as the Allen-Cahn or Cahn-Hilliard types, require careful numerical discretization to capture the diffuse interface dynamics accurately. Spatial discretization methods convert these continuum equations into algebraic systems suitable for computation, with common approaches including , finite element (particularly Galerkin-based), and methods. methods approximate derivatives on structured grids using expansions, offering simplicity and efficiency for regular domains, but they may struggle with complex geometries without modifications. Finite element methods, employing variational formulations, provide flexibility for irregular domains and adaptive refinement, where basis functions like piecewise polynomials ensure conformity and higher-order accuracy. methods, such as pseudospectral techniques, excel in periodic settings by leveraging global basis functions for exponential convergence, though they demand uniform grids and are less suited to non-periodic boundaries. To resolve the sharp gradients near interfaces, where the phase-field variable transitions over a width proportional to the parameter ε, adaptive mesh refinement is essential, dynamically increasing resolution in interface regions while coarsening elsewhere to optimize computational cost. These adaptive strategies often use error indicators based on gradient estimates or residual norms to guide mesh adaptation, achieving second-order convergence in space for sufficiently refined grids. For instance, in finite element implementations, hierarchical refinement allows local adjustments, ensuring the interface is spanned by multiple elements (typically 4–10) to maintain physical fidelity without excessive degrees of freedom. Temporal integration schemes address the evolution of the discretized system over time, balancing accuracy, , and given the arising from small ε. Explicit methods, such as Runge-Kutta schemes of orders 2–4, advance solutions directly but are conditionally stable, requiring time steps Δt satisfying a Courant-Friedrichs-Lewy (CFL) condition like Δt ≲ h²/ε, where h is the spatial mesh size, to prevent oscillations. Implicit schemes, including backward Euler or Crank-Nicolson, offer unconditional for linear problems and are preferred for stiff systems, solving nonlinear systems via Newton iterations at each step, though at higher per-step cost. For coupled phase-field models involving multiple fields (e.g., concentration and order parameter), operator splitting decomposes the evolution into substeps treating , reaction, and separately, enabling efficient explicit treatment of non-stiff terms while implicitly handling stiff ones, with preserved through convex splitting techniques. Numerical stability in phase-field simulations hinges on preserving key physical properties, such as energy dissipation, where the free energy functional decreases monotonically over time. Schemes that maintain this dissipation law at the discrete level, often via implicit-explicit (IMEX) hybrid approaches, avoid unphysical growth and ensure long-term reliability, particularly for small ε where explicit methods fail due to severe time-step restrictions. CFL conditions enforce stability by limiting Δt relative to the diffusive and interfacial timescales, with violations leading to numerical artifacts like interface pinning or spurious phase separation. Handling small ε demands either ε-independent schemes, such as adaptive time stepping that scales Δt ~ ε, or stabilized formulations to mitigate the O(1/ε) stiffness without grid refinement proportional to ε. Energy-stable discretizations, analyzed through discrete Lyapunov functionals, confirm second-order convergence in time for appropriately chosen parameters. For large-scale phase-field simulations, parallelization via domain decomposition partitions the spatial domain into subdomains assigned to processors, facilitating distributed computation of local contributions before global assembly. Overlapping Schwarz methods or non-overlapping FETI approaches handle inter-domain coupling, achieving near-linear speedup on thousands of cores for problems, with communication overhead minimized through exchanges. Error analysis for these methods reveals optimal rates: first- or second-order in space and time, with error bounds O(h^k + Δt^m + ε), where k and m depend on the scheme order, verified through a priori estimates and numerical benchmarks showing grid-independent accuracy for ε ≳ h. Parallel implementations must also control load imbalance from adaptive meshes, using dynamic repartitioning to sustain .

Software Tools and Frameworks

Several open-source software tools have been developed to facilitate phase-field simulations, particularly for multi-phase microstructures and related phenomena. OpenPhase is a modular C++-based open-source framework initiated in 2008 at , designed for versatile multi-physics simulations including solidification, , and phase transformations. It supports through integration with MPI for parallel execution and includes built-in modules for and validation against experimental data, enabling rapid workflow development for complex microstructural evolutions. Similarly, FiPy, developed by NIST, is a Python-based finite volume solver that implements phase-field models such as the Allen-Cahn equation for polycrystalline and dendritic transformations. Its object-oriented structure allows for easy customization, with examples for electrochemical phase changes and drug-eluting stents, alongside tools for and result . Another prominent tool is PRISMS-PF from the PRISMS Center, a massively parallel finite element code emphasizing performance and flexibility for microstructural evolution simulations, including crystal plasticity couplings. It features matrix-free solvers for efficiency, validation benchmarks, and MPI support for large-scale computations on . Commercial and academic frameworks also provide robust options for phase-field modeling with user-friendly interfaces. offers a dedicated Phase Field Model node within its CFD and Multiphysics modules, enabling graphical setup for couplings like , , and solid-state transformations such as . The software includes built-in visualization tools, parametric studies, and validation examples for three-phase flow and interface tracking, though it requires licensing for full access. The framework, an open-source finite element multiphysics environment from , incorporates a Phase Field Module for solving Cahn-Hilliard and Allen-Cahn equations in multiphase systems. It supports advanced features like split-operator solvers, via PETSc and MPI, and integration with visualization software such as , making it suitable for large-scale simulations of growth and . Recent trends in phase-field software emphasize acceleration through , particularly neural network-based surrogate models to expedite microstructure predictions in the . These approaches replace computationally intensive phase-field steps with data-driven approximations, achieving speedups while maintaining accuracy for applications like alloy design and materials. Frameworks like and PRISMS-PF are increasingly incorporating such integrations for hybrid simulations on platforms.

References

  1. [1]
    [PDF] Phase-field models in materials science
    Jul 30, 2009 · Abstract. The phase-field method is reviewed against its historical and theoretical background. Starting from Van der Waals considerations ...
  2. [2]
    [PDF] Computational phase-field modeling - Purdue Engineering
    Phase-field modeling is emerging as a promising tool for the treatment of problems with interfaces. The classical description of interface problems requires ...
  3. [3]
    [PDF] From classical thermodynamics to phase-field method - OSTI.GOV
    Phase-field method is a mesoscale computational technique for modeling and predicting spatial and temporal evolution of materials microstructures/domain ...
  4. [4]
    Free Energy of a Nonuniform System. I. Interfacial ... - AIP Publishing
    Cahn , John E. Hilliard. General Electric Research Laboratory, Schenectady, New York. J. Chem. Phys. 28, 258–267 (1958). https://doi.org/10.1063/1.1744102.Missing: equation | Show results with:equation
  5. [5]
    A microscopic theory for antiphase boundary motion and its ...
    A microscopic diffusional theory for the motion of a curved antiphase boundary is presented. The interfacial velocity is found to be linearly proportional ...
  6. [6]
    Phase-Field Simulation of Solidification - Annual Reviews
    Aug 1, 2002 · Abstract. An overview of the phase-field method for modeling solidification is presented, together with several example results.
  7. [7]
    Concepts of modeling surface energy anisotropy in phase-field ...
    Oct 11, 2017 · To simulate the growth of geological veins, it is necessary to model the crystal shape anisotropy. Two different models, classical and ...Background · Phase-Field Model · Multiphase-Field Models
  8. [8]
    Formulation of strongly non-local, non-isothermal dynamics for ...
    May 25, 2017 · The purpose of the current work is the formulation of models for conservative and non-conservative dynamics in solid systems with the help ...
  9. [9]
    [PDF] Phase-Field Methods in Material Science and Engineering
    free energy, which hopefully pushes the order parameter to minimum of the free energy landscape. The following subsections outline the basic evolution ...Missing: seminal | Show results with:seminal
  10. [10]
    [PDF] Phase-field models
    Feb 28, 2014 · Phase-field models are a mathematical tool for modeling crystal growth, using out-of-equilibrium thermodynamics and order parameters to ...Missing: seminal | Show results with:seminal
  11. [11]
    Thin interface asymptotics for an energy/entropy approach to phase ...
    Sep 15, 2000 · Almgren shows that a continuous temperature field and a Gibbs ... The thin interface asymptotics of Karma and Rappel applied to the standard phase ...
  12. [12]
    Phase-field models in materials science - IOPscience
    Jul 30, 2009 · The phase-field method is reviewed against its historical and theoretical background. Starting from Van der Waals considerations on the structure of interfaces ...
  13. [13]
    Mean curvature, threshold dynamics, and phase field theory on finite ...
    Jun 28, 2013 · Abstract page for arXiv paper 1307.0045: Mean curvature, threshold dynamics, and phase field theory on finite graphs. ... discrete time graph mean ...
  14. [14]
    Mean Curvature, Threshold Dynamics, and Phase Field Theory on ...
    Apr 18, 2014 · The aim of this paper is to introduce these graph processes in the light of their continuum counterparts, provide some background, prove the first results ...
  15. [15]
    A phase-field lattice model (PFLM) for fracture problem: Theory and ...
    Nov 1, 2023 · A three-dimensional phase-field lattice model is proposed for fracture problems. Failures of homogeneous and heterogeneous structures can be easily modelled ...2. Unified Phase-Field... · 3. Phase-Field Lattice Model · 5. Verification Of Numerical...
  16. [16]
    Phase-field method for computationally efficient modeling of ...
    Abstract. We present mathematical results which dramatically enhance the computational efficiency of the phase-field method for modeling the solidification of a ...<|control11|><|separator|>
  17. [17]
    Phase-field model for solidification of a eutectic alloy - Journals
    In this paper we discuss two phase-field models for solidification of a eutectic alloy, a situation in which a liquid may transform into two distinct solid ...
  18. [18]
    Dendritic solidification of Succinonitrile-0.24 wt% water alloy
    This study aims to provide an experimental benchmark of dendritic growth applicable for validation of theoretical and numerical dendrite growth models.
  19. [19]
    On/off directional solidification of near peritectic TRIS-NPG with a ...
    Here, we report about a single directional solidification experiment using a TRIS-NPG alloy with C0 = 53 mol% NPG and a pulling velocity that ensured a ...Missing: validation | Show results with:validation
  20. [20]
    A phase field model for rate-independent crack propagation
    In the recent work Miehe, Welschinger & Hofacker [1], we outlined a general thermodynamically-consistent framework of phase field modeling in fracture mechanics ...Missing: seminal | Show results with:seminal
  21. [21]
    Phase-field modelling of ductile fracture: a variational gradient ...
    Apr 28, 2016 · This work outlines a novel variational-based theory for the phase-field modelling of ductile fracture in elastic–plastic solids undergoing large strains.Introduction · Introduction of primary field... · Variational phase-field... · ConclusionMissing: seminal | Show results with:seminal
  22. [22]
    A phase-field description of dynamic brittle fracture - ScienceDirect
    In this work, we extend a phase-field model for quasi-static brittle fracture to the dynamic case. We introduce a phase-field approximation to the Lagrangian ...
  23. [23]
    A review on phase field models for fracture and fatigue - ScienceDirect
    Sep 1, 2023 · The purpose of this paper is to provide a comprehensive review of recent work on phase field models for fatigue damage and failures.
  24. [24]
    Full spectrum of grain boundary energy landscape in phase-field ...
    In this study, we examine grain boundary energy as a function of both misorientation and inclination using a phase-field approach.Abstract · Introduction · Methodology and validation · Results and discussions
  25. [25]
    Quantitative analysis of grain boundary properties in a generalized ...
    Jul 16, 2008 · In this paper, it is studied how the model parameters of a generalized phase field model affect the landscape of the free-energy density functional.
  26. [26]
    Phase coarsening in multicomponent systems | Phys. Rev. E
    Feb 27, 2017 · Lifshitz and Slyozov' theory in binary systems. The theoretical study of three-dimensional (3D) binary phase coarsening perhaps began in earnest ...Missing: impurities | Show results with:impurities
  27. [27]
    Ultra-large-scale phase-field simulation study of ideal grain growth
    Jul 3, 2017 · We analyze ideal grain growth via ultra-large-scale phase-field simulations on a supercomputer for elucidating the corresponding authentic statistical ...
  28. [28]
    Phase-field Simulations of Grain Growth in Two-dimensional ...
    Aug 10, 2025 · In this study, we have performed phase-field simulations of grain growth in two-dimensional systems containing finely dispersed coarsening ...
  29. [29]
    A coupled coarse-grained dislocation density and phase-field ...
    Sep 14, 2007 · In this work, we develop a coupled dislocation density and phase-field method to model the isothermal recrystallization process as a phase ...
  30. [30]
    Computationally Efficient Algorithm for Modeling Grain Growth Using ...
    May 15, 2024 · Grain growth can occur in different regimes, referred to as “normal” and “abnormal” growth. Normal grain growth is typically characterized by ...
  31. [31]
    Phase field modeling of microstructure evolution in steels
    This article provides an overview on the application of phase field models to describe microstructure evolution in steels. The focus will be on phase field ...
  32. [32]
    Phase field modeling of grain growth in thin films on rigid substrates
    Nov 2, 2011 · In phase field models, grain boundaries are approximated by diffuse interfaces, where the texture and in-plane orientation of grains change ...
  33. [33]
  34. [34]
  35. [35]
  36. [36]
  37. [37]
    Phase field modeling of electrochemistry. II. Kinetics | Phys. Rev. E
    Feb 27, 2004 · The kinetic behavior of a phase field model of electrochemistry is explored for advancing (electrodeposition) and receding (electrodissolution) conditions in ...
  38. [38]
    A phase-field model for electrode reactions with Butler–Volmer kinetics
    A novel phase-field model for electrochemical processes is formulated, in which the Butler–Volmer kinetics at the electrode/electrolyte interface is taken ...
  39. [39]
    [PDF] Quantitative Phase Field Model for Electrochemical Systems
    Oct 18, 2023 · A higher-order kinetic correction is derived to accurately reproduce general reaction models such as the. Butler-Volmer, Marcus, and Marcus-Hush ...
  40. [40]
    [PDF] Nonlinear phase-field model for electrode-electrolyte interface ...
    Nov 26, 2012 · It produces Butler-Volmer-type electrochemical kinetics for the dependence of interfacial velocity on the overpotential at the sharp ...
  41. [41]
    Phase-Field Investigation of Lithium Electrodeposition at Different ...
    Mar 28, 2022 · The phase-field method (32,33) is one of the most popular computational methods for modeling dendrite growth, as it is well-suited to simulate ...Introduction · Methodology · Results and Discussion · References
  42. [42]
    Phase-field study of dendritic morphology in lithium metal batteries
    Feb 1, 2021 · This study presents a phase-field method for modeling the lithium dendrite formation in LMBs. In this model, the electrochemical kinetics is described.
  43. [43]
    Quantitative phase-field modeling of dendritic electrodeposition
    Jul 6, 2015 · Abstract. A thin-interface phase-field model of electrochemical interfaces is developed based on Marcus kinetics for concentrated solutions, and ...Missing: seminal | Show results with:seminal
  44. [44]
    Phase-field model of pitting corrosion kinetics in metallic materials
    Jul 24, 2018 · This study presents a thermodynamically consistent phase field model for the quantitative prediction of the pitting corrosion kinetics in metallic materials.
  45. [45]
    Phase-field modeling for pH-dependent general and pitting ... - NIH
    This study proposes a new phase-field (PF) model to simulate the pH-dependent corrosion of iron. The model is formulated based on Bockris's iron dissolution ...
  46. [46]
    A phase field model for simulating the pitting corrosion - ScienceDirect
    This manuscript presents the formulation and implementation of a phase field model for simulating the activation- and diffusion-controlled pitting corrosion ...
  47. [47]
    Numerical Prediction of the Secondary Dendrite Arm Spacing Using ...
    The secondary arm spacing in Fe–C, Fe–P, Fe–C–P and Al–Cu alloys are numerically predicted using a phase-field model. The calculated arm spacing and the ...
  48. [48]
    Prediction of Secondary‐Dendrite Arm Spacing for Binary Alloys by ...
    The phase-field models were developed mainly for studying solidification of pure materials, being subsequently extended to the solidification of binary, ternary ...
  49. [49]
    Application of phase-field method in rechargeable batteries - Nature
    Nov 19, 2020 · In 2003, Monroe and Newman first established an electrochemical model for dendrite growth, which is aimed at evolution of dendrite tip ...
  50. [50]
    Numerical Implementation — Phase Field Method Recommended ...
    Four main types of spatial discretizations are used for phase-field modeling: finite difference, finite volume, finite element, and Fourier pseudospectral.
  51. [51]
    The phase field method for geometric moving interfaces and their ...
    Spatial discretization of phase field models has been widely studied, ranging from conventional finite difference, finite element and spectral methods, to ...
  52. [52]
    Conservative phase-field method with a parallel and adaptive-mesh ...
    Aug 14, 2019 · To achieve large-scale interface tracking, a parallel and adaptive-mesh-refinement algorithm is developed to reduce the computing overhead.
  53. [53]
    A spatially adaptive phase-field model of fracture - ScienceDirect.com
    This extended phase-field model serves as the foundation for an adaptive mesh refinement strategy, in which the mesh size is determined locally by the optimal ...
  54. [54]
    [PDF] On effective numerical methods for phase-field models
    The first part of the article will discuss implicit-explicit time discretizations which satisfy the energy stability. The second part is to discuss time- ...Missing: temporal | Show results with:temporal
  55. [55]
    Fast and stable explicit operator splitting methods for phase-field ...
    In this paper, we propose fast and stable explicit operator splitting methods for both one- and two-dimensional nonlinear diffusion equations for thin film ...Missing: temporal | Show results with:temporal
  56. [56]
    Numerical Energy Dissipation for Time-Fractional Phase-Field ...
    Sep 14, 2020 · In this article, we study certain nonlocal-in-time energies using the first-order stabilized semi-implicit L1 scheme.Missing: stability CFL condition small epsilon
  57. [57]
    Massively parallel phase field fracture simulations on supercomputers
    Dec 24, 2024 · The parallel implementation of the phase field method is based on domain decomposition, each process performs computations on a subdomain.Phase Field Method For... · Parallel Solvers And... · Notched Homogeneous...
  58. [58]
    Domain decomposition methods and acceleration techniques for the ...
    Jun 5, 2024 · This study proposes a domain decomposition framework and acceleration techniques for the phase field fracture staggered solver.
  59. [59]
    OpenPhase | Academic
    A versatile multi-physics open source software by combining the phase-field approach, well known to deal with solidification, grain growth and phase ...Citing OpenPhase · Download · About · Documentation
  60. [60]
    Tutorial 1: OpenPhase - SpringerLink
    Mar 23, 2023 · OpenPhase is a robust open-source microstructure simulation software project targeted at the phase-field simulations of complex scientific problems.
  61. [61]
    FiPy: A Finite Volume PDE Solver Using Python - NIST Pages
    FiPy is an object oriented, partial differential equation (PDE) solver, written in Python, based on a standard finite volume (FV) approach.Installation · Using FiPy · Python Module Index · Frequently Asked Questions
  62. [62]
    examples.phase.simple — FiPy 4.0.1+2.g8a16c2158 documentation
    Sep 30, 2025 · Solve a phase-field (Allen-Cahn) problem in one-dimension. To run this example from the base FiPy directory, type python examples/phase/simple/ ...
  63. [63]
    PRISMS-PF - GitHub Pages
    PRISMS-PF is a powerful, massively parallel finite element code for conducting phase field and other related simulations of microstructural evolution.
  64. [64]
    PRISMS-PF: A general framework for phase-field modeling with a ...
    Mar 26, 2020 · A new phase-field modeling framework with an emphasis on performance, flexibility, and ease of use is presented.
  65. [65]
    Phase Field Model
    The Phase Field Model node adds the equations described in The Equations for the Phase Field Method. The node defines the associated phase field parameters.
  66. [66]
    Simulate Three-Phase Flow with a New Phase Field Interface
    Dec 16, 2015 · In COMSOL Multiphysics version 5.2, the CFD and Microfluidics modules include a new fluid flow interface for modeling separated three-phase flow.
  67. [67]
    Phase Field Module - MOOSE framework
    MOOSE provides capabilities that enable the easy development of multiphase field model. A free energy expression has to be provided for each individual phase.Basic Phase Field EquationsNewton solverPhase Field ModuleIntroductionPhase Field System Design ...
  68. [68]
    Compute Linear Elastic Phase Field Fracture Stress | MOOSE
    This material implements the unified phase-field model for mechanics of damage and quasi-brittle failure from Jian-Ying Wu Wu (2017).
  69. [69]
    Accelerating microstructure modeling via machine learning
    Aug 7, 2023 · Replacing phase-field steps with machine-learned microstructures can significantly accelerate the in silico study of microstructure evolution.Missing: 2020s | Show results with:2020s
  70. [70]
    Accelerating phase-field simulation of coupled microstructural ...
    Jun 24, 2025 · Accelerated phase-field frameworks leveraging time-dependent neural networks have recently been developed to accelerate microstructure-based ...Missing: 2020s | Show results with:2020s