Phase field models

Last updated

A phase field model is a mathematical model for solving interfacial problems. It has mainly been applied to solidification dynamics, [1] but it has also been applied to other situations such as viscous fingering, [2] fracture mechanics, [3] [4] [5] [6] , hydrogen embrittlement [7] , and vesicle dynamics. [8]

Contents

The method substitutes boundary conditions at the interface by a partial differential equation for the evolution of an auxiliary field (the phase field) that takes the role of an order parameter. This phase field takes two distinct values (for instance +1 and 1) in each of the phases, with a smooth change between both values in the zone around the interface, which is then diffuse with a finite width. A discrete location of the interface may be defined as the collection of all points where the phase field takes a certain value (e.g., 0).

A phase field model is usually constructed in such a way that in the limit of an infinitesimal interface width (the so-called sharp interface limit) the correct interfacial dynamics are recovered. This approach permits to solve the problem by integrating a set of partial differential equations for the whole system, thus avoiding the explicit treatment of the boundary conditions at the interface.

Phase field models were first introduced by Fix [9] and Langer, [10] and have experienced a growing interest in solidification and other areas.

Equations of the Phase field model

Phase field models are usually constructed in order to reproduce a given interfacial dynamics. For instance, in solidification problems the front dynamics is given by a diffusion equation for either concentration or temperature in the bulk and some boundary conditions at the interface (a local equilibrium condition and a conservation law), [11] which constitutes the sharp interface model.

A two phase microstructure and the order parameter
ph
{\displaystyle \phi }
profile is shown on a line across the domain. Gradual change of order parameter from one phase to another shows diffuse nature of the interface. Phase field order parameter.jpg
A two phase microstructure and the order parameter profile is shown on a line across the domain. Gradual change of order parameter from one phase to another shows diffuse nature of the interface.

A number of formulations of the phase field model are based on a free energy function depending on an order parameter (the phase field) and a diffusive field (variational formulations). Equations of the model are then obtained by using general relations of statistical physics. Such a function is constructed from physical considerations, but contains a parameter or combination of parameters related to the interface width. Parameters of the model are then chosen by studying the limit of the model with this width going to zero, in such a way that one can identify this limit with the intended sharp interface model.

Other formulations start by writing directly the phase field equations, without referring to any thermodynamical functional (non-variational formulations). In this case the only reference is the sharp interface model, in the sense that it should be recovered when performing the small interface width limit of the phase field model.

Phase field equations in principle reproduce the interfacial dynamics when the interface width is small compared with the smallest length scale in the problem. In solidification this scale is the capillary length , which is a microscopic scale. From a computational point of view integration of partial differential equations resolving such a small scale is prohibitive. However, Karma and Rappel introduced the thin interface limit, [12] which permitted to relax this condition and has opened the way to practical quantitative simulations with phase field models. With the increasing power of computers and the theoretical progress in phase field modelling, phase field models have become a useful tool for the numerical simulation of interfacial problems.

Variational formulations

A model for a phase field can be constructed by physical arguments if one has an explicit expression for the free energy of the system. A simple example for solidification problems is the following:

where is the phase field, , is the local enthalpy per unit volume, is a certain polynomial function of , and (where is the latent heat, is the melting temperature, and is the specific heat). The term with corresponds to the interfacial energy. The function is usually taken as a double-well potential describing the free energy density of the bulk of each phase, which themselves correspond to the two minima of the function . The constants and have respectively dimensions of energy per unit length and energy per unit volume. The interface width is then given by . The phase field model can then be obtained from the following variational relations: [13]

where D is a diffusion coefficient for the variable , and and are stochastic terms accounting for thermal fluctuations (and whose statistical properties can be obtained from the fluctuation dissipation theorem). The first equation gives an equation for the evolution of the phase field, whereas the second one is a diffusion equation, which usually is rewritten for the temperature or for the concentration (in the case of an alloy). These equations are, scaling space with and times with :

where is the nondimensional interface width, , and , are nondimensionalized noises.

Alternative Energy Density Functions

The choice of free energy function, , can have a significant effect on the physical behaviour of the interface, and should be selected with care. The double-well function represents an approximation of the Van der Waals EOS near the critical point, and has historically been used for its simplicity of implementation when the phase field model is employed solely for interface tracking purposes. However, this has led to the frequently observed spontaneous drop shrinkage phenomenon, whereby the high phase miscibility predicted by an Equation of State near the critical point allows significant interpenetration of the phases and can eventually lead to the complete disappearance of a droplet whose radius is below some critical value. [14] Minimizing perceived continuity losses over the duration of a simulation requires limits on the Mobility parameter, resulting in a delicate balance between interfacial smearing due to convection, interfacial reconstruction due to free energy minimization (i.e. mobility-based diffusion), and phase interpenetration, also dependent on the mobility. A recent review of alternative energy density functions for interface tracking applications has proposed a modified form of the double-obstacle function which avoids the spontaneous drop shrinkage phenomena and limits on mobility, [15] with comparative results provide for a number of benchmark simulations using the double-well function and the VOF sharp interface technique. The proposed implementation has a computational complexity only slightly greater than that of the double-well function, and may prove useful for interface tracking applications of the phase field model where the duration/nature of the simulated phenomena introduces phase continuity concerns (i.e. small droplets, extended simulations, multiple interfaces, etc.).

Sharp interface limit of the Phase field equations

A phase field model can be constructed to purposely reproduce a given interfacial dynamics as represented by a sharp interface model. In such a case the sharp interface limit (i.e. the limit when the interface width goes to zero) of the proposed set of phase field equations should be performed. This limit is usually taken by asymptotic expansions of the fields of the model in powers of the interface width . These expansions are performed both in the interfacial region (inner expansion) and in the bulk (outer expansion), and then are asymptotically matched order by order. The result gives a partial differential equation for the diffusive field and a series of boundary conditions at the interface, which should correspond to the sharp interface model and whose comparison with it provides the values of the parameters of the phase field model.

Whereas such expansions were in early phase field models performed up to the lower order in only, more recent models use higher order asymptotics (thin interface limits) in order to cancel undesired spurious effects or to include new physics in the model. For example, this technique has permitted to cancel kinetic effects, [12] to treat cases with unequal diffusivities in the phases, [16] to model viscous fingering [2] and two-phase Navier–Stokes flows, [17] to include fluctuations in the model, [18] etc.

Multi Phase Field Models

Multiple order parameters describe a polycrystalline material microstructure. Multi Phase Field Order Parameters.jpg
Multiple order parameters describe a polycrystalline material microstructure.

In multi-phase field models, microstructure is described by set of order parameters, each of which is related to a specific phase or crystallographic orientation. This model is mostly used for solid state phase transformations where multiple grains evolve (e.g. grain growth, recrystallization or first order transformation like austenite to ferrite in ferrous alloys). Besides allowing the description of multiple grains in a microstructure, multi-phase field models especially allow for consideration of multiple thermodynamic phases occurring e.g. in technical alloy grades. [19]

Phase field models on graphs

Many of the results for continuum phase field models have discrete analogues for graphs, just replacing calculus with calculus on graphs.

Software

Related Research Articles

In mathematics, Poisson's equation is a partial differential equation of elliptic type with broad utility in mechanical engineering and theoretical physics. It arises, for instance, to describe the potential field caused by a given charge or mass density distribution; with the potential field known, one can then calculate gravitational or electrostatic field. It is a generalization of Laplace's equation, which is also frequently seen in physics. The equation is named after the French mathematician, geometer, and physicist Siméon Denis Poisson.

The Klein–Gordon equation is a relativistic wave equation, related to the Schrödinger equation. It is second-order in space and time and manifestly Lorentz-covariant. It is a quantized version of the relativistic energy–momentum relation. Its solutions include a quantum scalar or pseudoscalar field, a field whose quanta are spinless particles. Its theoretical relevance is similar to that of the Dirac equation. Electromagnetic interactions can be incorporated, forming the topic of scalar electrodynamics, but because common spinless particles like the pions are unstable and also experience the strong interaction the practical utility is limited.

In physics, chemistry and biology, a potential gradient is the local rate of change of the potential with respect to displacement, i.e. spatial derivative, or gradient. This quantity frequently occurs in equations of physical processes because it leads to some form of flux.

In physics, mathematics and statistics, scale invariance is a feature of objects or laws that do not change if scales of length, energy, or other variables, are multiplied by a common factor, and thus represent a universality.

Magnetic potential Integral of the magnetic field

The term magnetic potential can be used for either of two quantities in classical electromagnetism: the magnetic vector potential, or simply vector potential, A; and the magnetic scalar potentialψ. Both quantities can be used in certain circumstances to calculate the magnetic field B.

Proca action

In physics, specifically field theory and particle physics, the Proca action describes a massive spin-1 field of mass m in Minkowski spacetime. The corresponding equation is a relativistic wave equation called the Proca equation. The Proca action and equation are named after Romanian physicist Alexandru Proca.

A classical field theory is a physical theory that predicts how one or more physical fields interact with matter through field equations. The term 'classical field theory' is commonly reserved for describing those physical theories that describe electromagnetism and gravitation, two of the fundamental forces of nature. Theories that incorporate quantum mechanics are called quantum field theories.

Magnetostatics Branch of physics concerned with magnetic behavior in systems with steady electric currents

Magnetostatics is the study of magnetic fields in systems where the currents are steady. It is the magnetic analogue of electrostatics, where the charges are stationary. The magnetization need not be static; the equations of magnetostatics can be used to predict fast magnetic switching events that occur on time scales of nanoseconds or less. Magnetostatics is even a good approximation when the currents are not static — as long as the currents do not alternate rapidly. Magnetostatics is widely used in applications of micromagnetics such as models of magnetic storage devices as in computer memory. Magnetostatic focussing can be achieved either by a permanent magnet or by passing current through a coil of wire whose axis coincides with the beam axis.

In electromagnetism, the Lorenz gauge condition or Lorenz gauge is a partial gauge fixing of the electromagnetic vector potential. The condition is that This does not completely determine the gauge: one can still make a gauge transformation where is a harmonic scalar function.

Stokes flow

Stokes flow, also named creeping flow or creeping motion, is a type of fluid flow where advective inertial forces are small compared with viscous forces. The Reynolds number is low, i.e. . This is a typical situation in flows where the fluid velocities are very slow, the viscosities are very large, or the length-scales of the flow are very small. Creeping flow was first studied to understand lubrication. In nature this type of flow occurs in the swimming of microorganisms and sperm and the flow of lava. In technology, it occurs in paint, MEMS devices, and in the flow of viscous polymers generally.

In quantum chemistry, the potential energy surfaces are obtained within the adiabatic or Born–Oppenheimer approximation. This corresponds to a representation of the molecular wave function where the variables corresponding to the molecular geometry and the electronic degrees of freedom are separated. The non separable terms are due to the nuclear kinetic energy terms in the molecular Hamiltonian and are said to couple the potential energy surfaces. In the neighbourhood of an avoided crossing or conical intersection, these terms cannot be neglected. One therefore usually performs one unitary transformation from the adiabatic representation to the so-called diabatic representation in which the nuclear kinetic energy operator is diagonal. In this representation, the coupling is due to the electronic energy and is a scalar quantity which is significantly easier to estimate numerically.

The electromagnetic wave equation is a second-order partial differential equation that describes the propagation of electromagnetic waves through a medium or in a vacuum. It is a three-dimensional form of the wave equation. The homogeneous form of the equation, written in terms of either the electric field E or the magnetic field B, takes the form:

Quantum vortex

In physics, a quantum vortex represents a quantized flux circulation of some physical quantity. In most cases quantum vortices are a type of topological defect exhibited in superfluids and superconductors. The existence of quantum vortices was first predicted by Lars Onsager in 1949 in connection with superfluid helium. Onsager reasoned that quantisation of vorticity is a direct consequence of the existence of a superfluid order parameter as a spatially continuous wavefunction. Onsager also pointed out that quantum vortices describe the circulation of superfluid and conjectured that their excitations are responsible for superfluid phase transitions. These ideas of Onsager were further developed by Richard Feynman in 1955 and in 1957 were applied to describe the magnetic phase diagram of type-II superconductors by Alexei Alexeyevich Abrikosov. In 1935 Fritz London published a very closely related work on magnetic flux quantization in superconductors. London's fluxoid can also be viewed as a quantum vortex.

Spinodal decomposition

Spinodal decomposition is a mechanism for the rapid unmixing of a mixture of liquids or solids from one thermodynamic phase to form two coexisting phases. As an example, consider a hot mixture of water and an oil. At high temperatures the oil and the water may mix to form a single thermodynamic phase in which water molecules are surrounded by oil molecules and vice versa. The mixture is then suddenly cooled to a temperature at which thermodynamic equilibrium favours an oil-rich phase coexisting with a water-rich phase. Spinodal decomposition then occurs when the mixture is such that there is essentially no barrier to nucleation of the new oil-rich and water-rich phases. In other words, the oil and water molecules immediately start to cluster together into microscopic water-rich and oil-rich clusters throughout the liquid. These clusters then rapidly grow and coalesce until there is a single macroscopic oil-rich cluster, the oil-rich phase, and a single water-rich cluster, the water-rich phase.

In vector calculus, a topic in pure and applied mathematics, a poloidal–toroidal decomposition is a restricted form of the Helmholtz decomposition. It is often used in the spherical coordinates analysis of solenoidal vector fields, for example, magnetic fields and incompressible fluids.

The fractional Schrödinger equation is a fundamental equation of fractional quantum mechanics. It was discovered by Nick Laskin (1999) as a result of extending the Feynman path integral, from the Brownian-like to Lévy-like quantum mechanical paths. The term fractional Schrödinger equation was coined by Nick Laskin.

The Holstein–Herring method, also called the surface Integral method, or Smirnov's method is an effective means of getting the exchange energy splittings of asymptotically degenerate energy states in molecular systems. Although the exchange energy becomes elusive at large internuclear systems, it is of prominent importance in theories of molecular binding and magnetism. This splitting results from the symmetry under exchange of identical nuclei.

In physics, Berry connection and Berry curvature are related concepts which can be viewed, respectively, as a local gauge potential and gauge field associated with the Berry phase or geometric phase. These concepts were introduced by Michael Berry in a paper published in 1984 emphasizing how geometric phases provide a powerful unifying concept in several branches of classical and quantum physics.

Surface hopping is a mixed quantum-classical technique that incorporates quantum mechanical effects into molecular dynamics simulations. Traditional molecular dynamics assume the Born-Oppenheimer approximation, where the lighter electrons adjust instantaneously to the motion of the nuclei. Though the Born-Oppenheimer approximation is applicable to a wide range of problems, there are several applications, such as photoexcited dynamics, electron transfer, and surface chemistry where this approximation falls apart. Surface hopping partially incorporates the non-adiabatic effects by including excited adiabatic surfaces in the calculations, and allowing for 'hops' between these surfaces, subject to certain criteria.

In theoretical physics, Hamiltonian field theory is the field-theoretic analogue to classical Hamiltonian mechanics. It is a formalism in classical field theory alongside Lagrangian field theory, and has applications in quantum field theory also.

References

  1. Boettinger, W. J.; Warren, J. A.; Beckermann, C.; Karma, A. (2002). "Phase-Field Simulation of Solidification". Annual Review of Materials Research. 32: 163–194. doi:10.1146/annurev.matsci.32.101901.155803.
  2. 1 2 Folch, R.; Casademunt, J.; Hernández-Machado, A.; Ramírez-Piscina, L. (1999). "Phase-field model for Hele-Shaw flows with arbitrary viscosity contrast. II. Numerical study". Physical Review E. 60 (2): 1734–40. arXiv: cond-mat/9903173 . Bibcode:1999PhRvE..60.1734F. doi:10.1103/PhysRevE.60.1734. PMID   11969955.
  3. 1 2 Bourdin, B.; Francfort, G.A.; Marigo, J-J. (April 2000). "Numerical experiments in revisited brittle fracture". Journal of the Mechanics and Physics of Solids. 48 (4): 797–826. doi:10.1016/S0022-5096(99)00028-9.
  4. 1 2 Bourdin, Blaise (2007). "Numerical implementation of the variational formulation for quasi-static brittle fracture". Interfaces and Free Boundaries: 411–430. doi:10.4171/IFB/171. ISSN   1463-9963.
  5. 1 2 Bourdin, Blaise; Francfort, Gilles A.; Marigo, Jean-Jacques (April 2008). "The Variational Approach to Fracture". Journal of Elasticity. 91 (1–3): 5–148. doi:10.1007/s10659-007-9107-3. ISSN   0374-3535.
  6. Karma, Alain; Kessler, David; Levine, Herbert (2001). "Phase-Field Model of Mode III Dynamic Fracture". Physical Review Letters. 87 (4): 045501. arXiv: cond-mat/0105034 . Bibcode:2001PhRvL..87d5501K. doi:10.1103/PhysRevLett.87.045501. PMID   11461627.
  7. Martinez-Paneda, Emilio; Golahmar, Alireza; Niordson, Christian (2018). "A phase field formulation for hydrogen assisted cracking". Computer Methods in Applied Mechanics and Engineering. 342: 742–761. doi:10.1016/j.cma.2018.07.021.
  8. Biben, Thierry; Kassner, Klaus; Misbah, Chaouqi (2005). "Phase-field approach to three-dimensional vesicle dynamics". Physical Review E. 72 (4): 041921. Bibcode:2005PhRvE..72d1921B. doi:10.1103/PhysRevE.72.041921. PMID   16383434.
  9. G.J. Fix, in Free Boundary Problems: Theory and Applications, Ed. A. Fasano and M. Primicerio, p. 580, Pitman (Boston, 1983).
  10. Langer, J. S. (1986). "Models of Pattern Formation in First-Order Phase Transitions". Directions in Condensed Matter Physics. Directions in Condensed Matter Physics. Series: Series on Directions in Condensed Matter Physics. Series on Directions in Condensed Matter Physics. 1. Singapore: World Scientific. pp. 165–186. Bibcode:1986SDCMP...1..165L. doi:10.1142/9789814415309_0005. ISBN   978-9971-978-42-6.
  11. Langer, J. S. (1980). "Instabilities and pattern formation in crystal growth". Reviews of Modern Physics. 52 (1): 1–28. Bibcode:1980RvMP...52....1L. doi:10.1103/RevModPhys.52.1.
  12. 1 2 Karma, Alain; Rappel, Wouter-Jan (1998). "Quantitative phase-field modeling of dendritic growth in two and three dimensions". Physical Review E. 57 (4): 4323. Bibcode:1998PhRvE..57.4323K. doi:10.1103/PhysRevE.57.4323.
  13. Hohenberg, P.; Halperin, B. (1977). "Theory of dynamic critical phenomena". Reviews of Modern Physics. 49 (3): 435. Bibcode:1977RvMP...49..435H. doi:10.1103/RevModPhys.49.435.
  14. Yue, Pengtao; Zhou, Chunfeng; Feng, James J. (2007). "Spontaneous shrinkage of drops and mass conservation in phase-field simulations". Journal of Computational Physics. 223 (1): 1–9. Bibcode:2007JCoPh.223....1Y. CiteSeerX   10.1.1.583.2109 . doi:10.1016/j.jcp.2006.11.020.
  15. Donaldson, A.A.; Kirpalani, D.M.; MacChi, A. (2011). "Diffuse interface tracking of immiscible fluids: Improving phase continuity through free energy density selection". International Journal of Multiphase Flow. 37 (7): 777. doi:10.1016/j.ijmultiphaseflow.2011.02.002.
  16. McFadden, G.B.; Wheeler, A.A.; Anderson, D.M. (2000). "Thin interface asymptotics for an energy/entropy approach to phase-field models with unequal conductivities". Physica D: Nonlinear Phenomena. 144 (1–2): 154–168. Bibcode:2000PhyD..144..154M. doi:10.1016/S0167-2789(00)00064-6. hdl:2060/20000014455.
  17. Jacqmin, David (1999). "Calculation of Two-Phase Navier–Stokes Flows Using Phase-Field Modeling". Journal of Computational Physics. 155 (1): 96–127. Bibcode:1999JCoPh.155...96J. doi:10.1006/jcph.1999.6332.
  18. Benítez, R.; Ramírez-Piscina, L. (2005). "Sharp-interface projection of a fluctuating phase-field model". Physical Review E. 71 (6): 061603. arXiv: cond-mat/0409707 . Bibcode:2005PhRvE..71f1603B. doi:10.1103/PhysRevE.71.061603. PMID   16089744.
  19. Schmitz, G. J.; Böttger, B.; Eiken, J.; Apel, M.; Viardin, A.; Carré, A.; Laschet, G. (2011). "Phase-field based simulation of microstructure evolution in technical alloy grades". International Journal of Advances in Engineering Sciences and Applied Mathematics. 2 (4): 126. doi:10.1007/s12572-011-0026-y.

Further reading