Skip to content

Phase-field model validation

On this page, we focus on a number of validation cases of the phase-field model used to simulate liquid-solid phase transitions. For information on the convergence of the model with grid spacing, please refer to the convergence page. A more detailed overview of the setups considered here, including the initial conditions and any analytic solutions are provided in the examples pages on Stefan problems and flows coupled to phase change.

1-D melting from a hot boundary: phase-field parameters

We begin by investigating the model parameter \(A\) introduced by the phase-field method. As a reminder, the governing equation for the phase field \(\phi\) takes the form

\[ \partial_t \phi = D \nabla^2 \phi - \frac{D}{\varepsilon^2} \phi (1 - \phi) (1 - 2\phi + A(T - T_m)) , \]

where \(D=1.2/\mathcal{S}A Pe_T\) represents the diffusion coefficient of the phase field, and \(S\) captures the strength of the coupling between the temperature field and the phase field.

With that in mind, let us now consider how the value of \(A\) affects the results produced by the phase-field method. As a reminder, \(A=\varepsilon H\Delta T/\Gamma\), where \(\Gamma\) is related to the surface energy of the interface. Larger values of \(A\) therefore reduce the impact of the Gibbs-Thomson effect, where high local curvature increases the local melting temperature. In our 1-D example, the value of \(A\) should therefore play no significant effect.

First, we consider the position of the phase boundary over time. The analytic solution for the phase boundary height is

\[ h(t) = 2\Lambda \sqrt{(t+t_0)/Pe_T} , \]

where \(\Lambda = 0.62\) for \(\mathcal{S}=1\), and \(t_0\) is a value such that \(h(0)=0.1\). In the below figure, we plot the analytic solution against the position of \(\phi=0.5\) for the phase-field model with \(1\leq A \leq 100\).

As expected, the value of \(A\) appears to have little impact on the phase-field solution. Small differences between the simulations develop, but this is likely due to the initial condition of the phase-field, which is taken to be the equilibrium solution \(\phi=0.5(1 - \tanh ((x-0.1)/2\varepsilon))\). This initial condition may not perfectly represent the initial motion of the interface at early times.

Now we can take a closer look at the temperature profiles as we vary the parameters. Below we plot the temperature profile \(T(x,t)\) for a range of discrete times between \(t=0\) and \(t=100\).

The results at first seem almost indistinguishable from the analytic solution for each value of \(A\). However on close inspection, we can see that the temperature in the solid phase near the phase boundary is slightly above the melting temperature \(T=0\) for the cases \(A>1\). This does not appear to play a significant role in the development of the temperature field in the liquid, or the interface position, but this result should be kept in mind when proceeding with the other validation cases.

1-D Freezing of a supercooled melt

As we shall see, the stability of the phase-field model is more sensitive to the parameter choices for the case of supercooling. In this example, the solid phase grows out of a liquid that is cooled to below the melting temperature. An analytic solution can be derived for a semi-infinite domain, which we can compare to our finite domain when the interface is sufficiently far from the upper boundary.

For \(A = 1\) the one-dimensional growth of the solid phase matches the analytic solution well, even as the effect of the upper boundary becomes more significant. As seen below, both the temperature profiles and the interface position are consistent with the analytic solution.

However, for higher values of \(A\) the model breaks down, and the phase field \(\phi\) takes begins to take values away from \(0\) and \(1\) in the liquid layer. The heat fluxes also appear inconsistent at the phase boundary, with relatively large temperatures spreading throughout the domain.

The reason for the instability of the system is currently unclear, but given these results, it seems sensible to fix the model parameter \(A=1\) in simulations where the effect of supercooling may occur.

2-D axisymmetric melting

We now consider our first two-dimensional problem, where curved phase boundaries arise and so changing the parameter \(A\) will affect the results through the Gibbs-Thomson effect. In particular, we follow the example from appendix A.2 of Favier et al. (2019), where a disc of initial radius \(r=0.1\) melts into a liquid of initially uniform temperature. A video showcasing the evolution of the temperature field is shown below.

Favier et al. (2019) provide a semi-analytic solution for the case in which surface energy is zero and the melting temperature on the phase boundary remains equal to \(T_m\). Below, we reproduce the two figures from their appendix. We compare the phase-field model for various values of \(A\), keeping \(\mathcal{S}=1\).

The initial development is quite similar in all cases, since the curvature of the interface is not too high. As the simulation develops and the disc shrinks, the interface curvature increases, leading to faster melting in the cases where \(A\) is smaller. This matches the trend seen by Favier et al. (2019), and as \(A\) gets larger, we appear to approach the semi-analytic solution associated with no Gibbs-Thomson effect.

This trend is even clearer in the temperature profiles, where for cases with lower \(A\) we see the interface temperature drop significantly as the disc melts. Reassuringly, despite such variation in the solid and phase boundary, the temperature profile in the liquid phase remains very similar between all cases.

2-D Rayleigh-Bénard convection with a melting boundary

Now, we introduce an example where the evolution of the temperature field and the phase boundary are also coupled to the fluid flow. This again follows a validation example used by Favier et al. (2019), and details of the initial condition can be found on the examples page. The system consists of a fluid heated from below that underlies a solid phase. The evolution of the system is shown in the following video:

We vary the model parameter \(A\) to see the effect on the results, while keeping the resolution fixed with \(n_x=512\), \(n_x^r=1024\). As may be expected, the corners that emerge in the interface for \(A=10^3\) are somewhat smoothed out by setting \(A\) to a lower value. This highlights the importance of choosing \(A\) based on realistic values of the surface energy for each simulation.

In the above figure, we also compare the two implementations for imposing zero velocity in the solid phase. The simple immersed boundary method is used for the results shown in coloured lines, whereas the black dashed lines use the volume penalty method.

1-D multicomponent diffusive melting

This validation case considers the diffusive example from Martin and Kauffman (1977) described here. Below, we present the results obtained from a run of the phase-field model with \(Pe_T=10^3\), \(Pe_S=10^4\), \(\mathcal{S}=2.5\), \(A=1\), and \(\Lambda=0.4\). No-flux boundary conditions are applied to \(T\) and \(C\) at \(x=0\) and \(x=1\), and a resolution of \(512\) points is used for \(T\), whereas \(C\) and \(\phi\) are resolved on a grid of \(1024\) points. As a reminder, the governing equations for the multicomponent phase-field model are

\[ \partial_t T = \kappa_T \nabla^2 T + \mathcal{S} \partial_t \phi , \]
\[ \partial_t C = \kappa_S \left(\nabla^2 C - \frac{\boldsymbol{\nabla} \phi \cdot \boldsymbol{\nabla} C}{1 - \phi + \delta}\right) + \frac{C\partial_t \phi}{1 - \phi + \delta}, \]
\[ \partial_t \phi = D \nabla^2 \phi - \frac{D}{\varepsilon^2} \phi (1 - \phi) [1 - 2\phi + A(T +\Lambda C)] . \]

The simulation is initialised with the analytic solution given above at time \(t=1\), and an initial phase-field profile of

\[ \phi = \frac{1}{2}\left[ 1 + \tanh \left(\frac{x - h_0}{2\varepsilon} \right) \right] \]

where \(h_0 = 2\alpha /\sqrt{Pe_T}\) is the corresponding interface position.

First, we inspect the temperature profile over time. The profiles match well, and the temperature in the solid remains at the constant value predicted by the analytic solution.

Inspection of the salinity field reveals good agreement once again in the liquid, although the output of the model does not jump to zero in the solid. Salinity must be continuous at the phase boundary, which does not match the physical solution, but we already know that the salinity must be zero in the solid phase. The phase-field model is set up such that the correct boundary conditions are applied at the interface, making the solution in the liquid reliable.

As further evidence of the model's good agreement, we plot the salinity profiles such that the concentration is set to zero in the solid phase (\(\phi>0.5\)):

The evolution of the phase-field interface over time also recovers the analytic \(t^{1/2}\) expression near-perfectly: