This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A decoupled, stable, and linear FEM for a phase-field model of variable density two-phase incompressible surface flow

Yerbol Palzhanov    Alexander Zhiliakov    Annalisa Quaini    Maxim Olshanskii
(Department of Mathematics, University of Houston, Houston, Texas 77204
{ypalzhanov}{azhiliakov}{aquaini}{maolshanskiy}@uh.edu
)
Abstract

The paper considers a thermodynamically consistent phase-field model of a two-phase flow of incompressible viscous fluids. The model allows for a non-linear dependence of fluid density on the phase-field order parameter. Driven by applications in biomembrane studies, the model is written for tangential flows of fluids constrained to a surface and consists of (surface) Navier–Stokes–Cahn–Hilliard type equations. We apply an unfitted finite element method to discretize the system and introduce a fully discrete time-stepping scheme with the following properties: (i) the scheme decouples the fluid and phase-field equation solvers at each time step, (ii) the resulting two algebraic systems are linear, and (iii) the numerical solution satisfies the same stability bound as the solution of the original system under some restrictions on the discretization parameters. Numerical examples are provided to demonstrate the stability, accuracy, and overall efficiency of the approach. Our computational study of several two-phase surface flows reveals some interesting dependencies of flow statistics on the geometry.

Keywords: Two-phase flow; Navier–Stokes–Cahn–Hilliard system; Surface PDEs, TraceFEM, Bio-membranes, Kevin–Helmholtz instability; Rayleigh–Taylor instability

1 Introduction

Due to many applications in the physical and biological sciences, binary (two-phase) flows have attracted increasing interest in the past decades. The phase-field method has emerged as a powerful and effective technique for modeling multi-phase problems in general. This method describes the physical system using a set of field variables that are continuous across the interfacial regions separating the phases. The main reasons for the success of the phase-field methodology are the following: it replaces sharp interfaces by thin transition regions called diffuse interfaces, making front-tracking unnecessary, and it is based on rigorous mathematics. The classical diffuse-interface description of binary incompressible fluid with density-matched components is the so-called Model H [22], which is a Navier–Stokes–Cahn–Hilliard (NSCH) system. Several generalization to the case of variable density components have been presented [2, 4, 6, 7, 8, 14, 15, 29, 42]. Some of these models relax the incompressibility constraint to quasi-compressibility and not all models satisfy Galilean invariance, local mass conservation or thermodynamic consistency. In this paper, we focus on the generalization of Model H first presented in [2]. This model is thermodynamically consistent when the density of the mixture depends linearly on the phase-field variable (concentration or volume/surface fraction). We present an extension of the model in [2] that is thermodynamically consistent for a general monotone relation of density and phase-field variable.

Driven by applications in the analysis of lateral organization of plasma membranes (see [45, 46] for more context), we adopt this new model to simulate two-phase flow dynamics on arbitrary-shaped closed smooth surfaces. While there exist many computational studies of multi-component fluid flows in planar and volumetric domains (see, e.g., [19, 28, 31, 44, 47] for some recent works), the number of papers where NSCH systems are treated on manifolds is limited. This can be explained by the fact that solving equations numerically on general surfaces poses additional difficulties that are related to the discretization of tangential differential operators and the approximate recovery of (possibly) complex shapes. In [30], the authors resort to a stream function formulation, decouple the surface Navier–Stokes problem from the surface Cahn-Hilliad problem, and use a parametric finite element approach. Yang and Kim [43] experimented with a finite difference method on staggered marker-and-cell meshes for a NSCH system embedded in the mesh.

In the present paper, we study a geometrically unfitted finite element method for the simulation of two-phase incompressible flow on surfaces. Our approach builds on earlier work on a unfitted FEM for elliptic PDEs posed on surfaces [36] called TraceFEM. Unlike some other geometrically unfitted methods for surface PDEs, TraceFEM employs sharp surface representation. The surface can be defined implicitly and no knowledge of the surface parametrization is required. This method allows to solve for a scalar quantity or a vector field on a surface, for which a parametrization or triangulation is not required. TaceFEM has been extended to the surface Stokes problem in [32, 33] and the surface Cahn–Hilliard problem in [45]. One of the main advantages of TraceFEM is that it works well for PDEs posed on evolving surfaces, including cases with topological changes [27]. Although we will not benefit from this advantage in this paper (as the surfaces are fixed), it will become handy in a future modeling of membrane deformation and fusion for multicomponent lipid bilayers.

Although technical, the numerical analysis of diffuse-interface models for two-phase fluids with matching densities can be largely built on well-established analyses for the incompressible Navier–Stokes equations and Cahn–Hilliad equations alone. See, for example, [12] for the convergence study of a coupled implicit FE method, [25] for the analysis of a decoupled FE scheme, and [21] for the analysis of a linear and second-order in time FE method. The design of energy stable, efficient, and consistent discretizations of diffuse-interface models for two-phase fluids with non-matching densities turns out to be more challenging.

Following the ideas from [18] for the incompressible Navier–Stokes equations with variable density, in [40] Shen and Yang suggested a stable time-stepping method for a NSCH system with the momentum equation modified based on an inconsistent mass conservation equation. In [41], the same authors introduced a time-stepping scheme for the thermodynamically consistent model from [2]. This scheme (which is first order and linear, and it decouples fluid and phase equations) introduces some extra terms in the momentum equation and modifies the advection flux for the order parameter to achieve energy stability. In the same paper, the authors present the stability analysis for a semi-discrete case. In [48], the approach from [41] was extended with the application of a scalar auxiliary variable method [38] and the stability was studied for a finite difference scheme on uniform staggered grids. A fully discrete FE method for the model from [2] was analyzed in [17], subject to the coupled treatment of the phase and fluid equations and a modification of the momentum equation. In [13], this analysis is extended to a posteriori estimates and adaptivity. The question of building provably stable and computationally efficient methods for thermodynamically consistent discretization models of two-phase fluids with variable densities remains largely open.

The present study contributes to the field with a first order in time, linear, and decoupled FE scheme for our generalization of the model from [2]. One advantage of our scheme is that it neither modifies the momentum equation nor alters the advection velocity. We prove that our scheme is stable under a mild time-step restriction. Moreover, the analysis tracks the dependence of the estimates on ϵ\epsilon, the critical model parameter that defines the width of the transition region between phases. The fact that the model is posed on arbitrary-shaped closed smooth surfaces and the use of an unfitted finite element method for spatial discretization add some further technical difficulties to our analysis. However, apart from these extra technical details, the general line of arguments can be followed for the simplified to Euclidian setting or extended to other spatial discretization techniques.

Numerical analysis is even less developed for quasi-incompressible diffuse-interface models due to their additional complexity, e.g. a more subtle nonlinearity. Here, we would like to mention the analysis of a semi-discrete method (coupled, but resulting in linear systems at each time step) in [42].

The rest of the paper is organized as follows. In Sec. 2, we state our generalization of the model from [2] and its variational formulations. A decoupled linear FEM based on the application of TraceFEM to our model is presented in Sec. 3, while the analysis is reported in Sec. 4. In Sec. 5, we report several numerical results, including a convergence test and the simulation of the Kevin–Helmholtz and Rayleigh–Taylor instabilities. Sec. 6 collects a few concluding remarks.

2 Problem definition

On surface Γ\Gamma we consider a heterogeneous mixture of two species with surface fractions ci=Si/Sc_{i}=S_{i}/S, i=1,2i=1,2, where SiS_{i} are the surface area occupied by the components and SS is the surface area of Γ\Gamma. Since S=S1+S2S=S_{1}+S_{2}, we have c1+c2=1c_{1}+c_{2}=1. Let c1c_{1} be the representative surface fraction, i.e c=c1c=c_{1}. Moreover, let mim_{i} be the mass of component ii and mm is the total mass. Notice that density of the mixture can be expressed as ρ=mS=m1S1S1S+m2S2S2S\rho=\frac{m}{S}=\frac{m_{1}}{S_{1}}\frac{S_{1}}{S}+\frac{m_{2}}{S_{2}}\frac{S_{2}}{S}. Thus,

ρ=ρ(c)=ρ1c+ρ2(1c).\displaystyle\rho=\rho(c)=\rho_{1}c+\rho_{2}(1-c). (2.1)

Similarly, for the dynamic viscosity of the mixture we can write

η=η(c)=η1c+η2(1c),\displaystyle\eta=\eta(c)=\eta_{1}c+\eta_{2}(1-c),~{} (2.2)

where η1\eta_{1} and η2\eta_{2} are the dynamic viscosities of the two species.

In order to state the model posed on surface Γ\Gamma, we need some preliminary notation. Let 𝐏=𝐏(𝐱)𝐈𝐧(𝐱)𝐧(𝐱)T\mathbf{P}=\mathbf{P}(\mathbf{x})\coloneqq\mathbf{I}-\mathbf{n}(\mathbf{x})\mathbf{n}(\mathbf{x})^{T} for 𝐱Γ\mathbf{x}\in\Gamma be the orthogonal projection onto the tangent plane. For a scalar function p:Γp:\,\Gamma\to\mathbb{R} or a vector function 𝐮:Γ3\mathbf{u}:\,\Gamma\to\mathbb{R}^{3} we define pe:𝒪(Γ)p^{e}\,:\,\mathcal{O}(\Gamma)\to\mathbb{R}, 𝐮e:𝒪(Γ)3\mathbf{u}^{e}\,:\,\mathcal{O}(\Gamma)\to\mathbb{R}^{3}, extensions of pp and 𝐮\mathbf{u} from Γ\Gamma to its neighborhood 𝒪(Γ)\mathcal{O}(\Gamma). The surface gradient and covariant derivatives on Γ\Gamma are then defined as Γp=𝐏pe\nabla_{\Gamma}p=\mathbf{P}\nabla p^{e} and Γ𝐮𝐏𝐮e𝐏\nabla_{\Gamma}\mathbf{u}\coloneqq\mathbf{P}\nabla\mathbf{u}^{e}\mathbf{P}. These definitions are independent of a particular smooth extension of pp and 𝐮\mathbf{u} off Γ\Gamma. On Γ\Gamma we consider the surface rate-of-strain tensor [20] given by

Es(𝐮)12(Γ𝐮+(Γ𝐮)T).E_{s}(\mathbf{u})\coloneqq\frac{1}{2}(\nabla_{\Gamma}\mathbf{u}+(\nabla_{\Gamma}\mathbf{u})^{T}). (2.3)

The surface divergence operators for a vector 𝐠:Γ3\mathbf{g}:\Gamma\to\mathbb{R}^{3} and a tensor 𝐀:Γ3×3\mathbf{A}:\Gamma\to\mathbb{R}^{3\times 3} are defined as:

divΓ𝐠tr(Γ𝐠),divΓ𝐀(divΓ(𝐞1T𝐀),divΓ(𝐞2T𝐀),divΓ(𝐞3T𝐀))T,{\mathop{\,\rm div}}_{\Gamma}\mathbf{g}\coloneqq{\rm tr}(\nabla_{\Gamma}\mathbf{g}),\qquad{\mathop{\,\rm div}}_{\Gamma}\mathbf{A}\coloneqq\left({\mathop{\,\rm div}}_{\Gamma}(\mathbf{e}_{1}^{T}\mathbf{A}),\,{\mathop{\,\rm div}}_{\Gamma}(\mathbf{e}_{2}^{T}\mathbf{A}),\,{\mathop{\,\rm div}}_{\Gamma}(\mathbf{e}_{3}^{T}\mathbf{A})\right)^{T},

with 𝐞i\mathbf{e}_{i} the iith standard basis vector in 3\mathbb{R}^{3}.

The classical phase-field model to describe the flow of two immiscible, incompressible, and Newtonian fluids is the so-called Model H [22]. One of the fundamental assumptions for Model H is that the densities of both components are matching. Several extensions have been proposed to account for the case of non-matching densities; see Sec. 1. Here, we restrict our attention to a thermodynamically consistent generalization of Model H first presented in [2]. For surface based quantities the model reads:

ρt𝐮+ρ(Γ𝐮)𝐮𝐏divΓ(2ηEs(𝐮))+Γp\displaystyle\rho\partial_{t}\mathbf{u}+\rho(\nabla_{\Gamma}\mathbf{u})\mathbf{u}-\mathbf{P}{\mathop{\,\rm div}}_{\Gamma}(2\eta E_{s}(\mathbf{u}))+\nabla_{\Gamma}p =σγcΓμ+Mdρdc(Γ𝐮)Γμ+𝐟,\displaystyle=-\sigma_{\gamma}c\nabla_{\Gamma}\mu+{M\frac{d\rho}{dc}(\nabla_{\Gamma}\mathbf{u})\nabla_{\Gamma}\mu}+\mathbf{f}, (2.4)
divΓ𝐮\displaystyle{\mathop{\,\rm div}}_{\Gamma}\mathbf{u} =0,\displaystyle=0, (2.5)
tc+divΓ(c𝐮)divΓ(MΓμ)\displaystyle\partial_{t}c+{\mathop{\,\rm div}}_{\Gamma}(c\mathbf{u})-{\mathop{\,\rm div}}_{\Gamma}\left(M\nabla_{\Gamma}\mu\right) =0,\displaystyle=0, (2.6)
μ\displaystyle\mu =1ϵf0ϵΔΓc.\displaystyle=\frac{1}{\epsilon}f_{0}^{\prime}-\epsilon\Delta_{\Gamma}c. (2.7)

Here, 𝐮\mathbf{u} is the surface averaged tangential velocity 𝐮=c𝐮1+(1c)𝐮2\mathbf{u}=c\mathbf{u}_{1}+(1-c)\mathbf{u}_{2}, density is given by (2.1) and viscosity by (2.2), σγ\sigma_{\gamma} is line tension and μ\mu is the chemical potential defined in (2.7). Force vector 𝐟\mathbf{f}, with 𝐟𝐧=0\mathbf{f}\cdot\mathbf{n}=0, is given. In (2.6) MM is the mobility coefficient, while in (2.7) f0(c)f_{0}(c) is the specific free energy of a homogeneous phase and ϵ\epsilon is a parameter related to the thickness of the interface between the two phases. We recall that in order to have phase separation, f0f_{0} must be a non-convex function of cc. For a comprehensive discussion of the related Navier–Stokes and Cahn–Hilliard equations on surfaces we refer, for example, to [23, 10, 9].

Problem (2.4)–(2.7) has only one additional term with respect to Model H: the middle term at the right-hand side in eq. (2.4). Notice that this term vanishes in the case of matching densities since dρdc=(ρ1ρ2)\frac{d\rho}{dc}=(\rho_{1}-\rho_{2}). However, the term is crucial for thermodynamic consistency when the densities do not match and it can be interpreted as additional momentum flux due to diffusion of the components driven by the gradient of the chemical potential.

In practice, we are interested in more general relations than (2.1) between ρ\rho and cc, since depending on the choice of f0(c)f_{0}(c) (and because of numerical errors while computing) the order parameter may not be constrained in [0,1][0,1] and so ρ\rho and η\eta based on (2.1) may take physically meaningless (even negative) values. Since we do not see how to show the thermodynamic consistency of (2.4)–(2.7) for non-linear ρ(c)\rho(c), we propose a further modification. Without the loss of generality let ρ1ρ2\rho_{1}\geq\rho_{2}. For a general dependence of ρ\rho on cc, it is reasonable to assume that ρ\rho is a smooth monotonic function of cc, i.e. dρdc0\frac{d\rho}{dc}\geq 0 (for ρ1ρ2\rho_{1}\geq\rho_{2}), and so we can set

dρdc=θ2.\frac{d\rho}{dc}=\theta^{2}. (2.8)

Then, we replace (2.4) with

ρt𝐮+ρ(Γ𝐮)𝐮𝐏divΓ(2ηEs(𝐮))+Γp=σγcΓμ+Mθ(Γ(θ𝐮))Γμ+𝐟.\rho\partial_{t}\mathbf{u}+\rho(\nabla_{\Gamma}\mathbf{u})\mathbf{u}-\mathbf{P}{\mathop{\,\rm div}}_{\Gamma}(2\eta E_{s}(\mathbf{u}))+\nabla_{\Gamma}p=-\sigma_{\gamma}c\nabla_{\Gamma}\mu+{M\theta(\nabla_{\Gamma}(\theta\mathbf{u})\,)\nabla_{\Gamma}\mu}+\mathbf{f}. (2.9)

The updated model (2.9),(2.5)–(2.7) obviously coincides with (2.4)–(2.7) for ρ(c)\rho(c) from (2.1), but exhibits thermodynamic consistency for a general monotone ρ\rhocc relation as we show below. The consistency is preserved if MM is a non-negative function of cc rather than a constant coefficient. Thermodynamic consistent extensions of (2.4)–(2.7) for a generic smooth ρ(c)\rho(c) (no monotonicity assumption) were also considered in [1, 3] for the purpose of well-posedness analysis. Those extensions introduce yet more term(s) in the momentum equation, so for computational needs and numerical analysis purpose we opt for (2.9).

From now on, we will focus on problem (2.9), (2.5)–(2.7). This is a Navier–Stokes–Cahn–Hilliard (NSCH) system that needs to be supplemented with the definitions of mobility MM and free energy per unit area f0f_{0}. Following many of the existing analytic studies, as well as numerical studies, we assume mobility to be constant (i.e., independent of cc). As for f0f_{0}, a classical choice is the Ginzburg–Landau double-well potential

f0(c)=ξ4c2(1c)2,\displaystyle f_{0}(c)=\frac{\xi}{4}c^{2}(1-c)^{2}, (2.10)

where ξ\xi defines the barrier height, i.e. the local maximum at c=1/2c=1/2 [11]. We set ξ=1\xi=1 for the rest of the paper.

For the numerical method, we need a weak (integral) formulation. We define the spaces

𝐕T{𝐮H1(Γ)3|𝐮𝐧=0},E{𝐮𝐕T|Es(𝐮)=𝟎}.\mathbf{V}_{T}\coloneqq\{\,\mathbf{u}\in H^{1}(\Gamma)^{3}~{}|~{}\mathbf{u}\cdot\mathbf{n}=0\,\},\quad E\coloneqq\{\,\mathbf{u}\in\mathbf{V}_{T}~{}|~{}E_{s}(\mathbf{u})=\mathbf{0}\,\}. (2.11)

We define the Hilbert space 𝐕T0\mathbf{V}_{T}^{0} as an orthogonal complement of EE in 𝐕T\mathbf{V}_{T} (hence 𝐕T0𝐕T/E\mathbf{V}_{T}^{0}\sim\mathbf{V}_{T}/E), and recall the surface Korn’s inequality [23]:

𝐮H1(Γ)CKEs(𝐮),𝐮𝐕T0.\|\mathbf{u}\|_{H^{1}(\Gamma)}\leq C_{K}\|E_{s}(\mathbf{u})\|,\quad\forall\,\mathbf{u}\in\mathbf{V}_{T}^{0}. (2.12)

In (2.12) and later we use short notation \|\cdot\| for the L2L^{2}-norm on Γ\Gamma. Finally, we define L02(Γ){pL2(Γ)|Γp𝑑s=0}L_{0}^{2}(\Gamma)\coloneqq\{\,p\in L^{2}(\Gamma)~{}|~{}\int_{\Gamma}p\,ds=0\,\}. To devise the weak formulation, one multiplies eq. (2.9) by 𝐯𝐕T\mathbf{v}\in\mathbf{V}_{T}, eq. (2.5) by qL2(Γ)q\in L^{2}(\Gamma), eq. (2.6) by vH1(Γ)v\in H^{1}(\Gamma), and eq. (2.7) by gH1(Γ)g\in H^{1}(\Gamma) and integrates all the equations over Γ\Gamma. For eq. (LABEL:gracke-1m) and (2.6), one employs the integration by parts identity, which for a closed smooth surface Γ\Gamma reads:

ΓvdivΓ𝐠𝑑s=Γ𝐠Γvds+Γκv𝐠𝐧𝑑s,for𝐠H1(Γ)3,vH1(Γ),\int_{\Gamma}v\textrm{div}\ \!_{\Gamma}\mathbf{g}\,ds=-\int_{\Gamma}\mathbf{g}\cdot\nabla_{\Gamma}v\,ds+\int_{\Gamma}\kappa v\mathbf{g}\cdot\mathbf{n}\,ds,\quad\text{for}~{}\mathbf{g}\in H^{1}(\Gamma)^{3},\,v\in H^{1}(\Gamma), (2.13)

here κ\kappa is the sum of principle curvatures. Identity (2.13) is applied to the second term in (2.6) (i.e. 𝐠=c𝐮\mathbf{g}=c\mathbf{u}), which leads to no contribution from the curvature term since 𝐮\mathbf{u} is tangential, and to the third term in (2.6) (i.e. 𝐠=MΓμ\mathbf{g}=M\nabla_{\Gamma}\mu), which makes the curvature term vanish as well. For a similar reason (component-wise), the curvature term vanishes also when identity (2.13) is applied to the diffusion term in (2.9).

The weak (integral) formulation of the surface NSCH problem (2.9), (2.5)–(2.7) reads: Find (𝐮,p,c,μ)𝐕T×L02(Γ)×H1(Γ)×H1(Γ)(\mathbf{u},p,c,\mu)\in\mathbf{V}_{T}\times L_{0}^{2}(\Gamma)\times H^{1}(\Gamma)\times H^{1}(\Gamma) such that

Γ(ρt𝐮𝐯+ρ(Γ𝐮)𝐮𝐯+2ηEs(𝐮):Es(𝐯))dsΓpdivΓ𝐯ds=ΓσγcΓμ𝐯ds\displaystyle\int_{\Gamma}\left(\rho\partial_{t}\mathbf{u}\cdot\mathbf{v}+\rho(\nabla_{\Gamma}\mathbf{u})\mathbf{u}\cdot\mathbf{v}+2\eta E_{s}(\mathbf{u}):E_{s}(\mathbf{v})\right)\,ds-\int_{\Gamma}p\,{\mathop{\,\rm div}}_{\Gamma}\mathbf{v}\,ds=-\int_{\Gamma}\sigma_{\gamma}c\nabla_{\Gamma}\mu\cdot\mathbf{v}\,ds
+ΓM(Γ(θ𝐮))(Γμ)(θ𝐯)𝑑s+Γ𝐟𝐯𝑑s,\displaystyle\quad\quad+\int_{\Gamma}M(\nabla_{\Gamma}(\theta\mathbf{u}))(\nabla_{\Gamma}\mu)\cdot(\theta\mathbf{v})\,ds+\int_{\Gamma}\mathbf{f}\cdot\mathbf{v}\,ds, (2.14)
ΓqdivΓ𝐮ds=0,\displaystyle\int_{\Gamma}q\,{\mathop{\,\rm div}}_{\Gamma}\mathbf{u}\,ds=0, (2.15)
ΓtcvdsΓc𝐮Γvds+ΓMΓμΓvds=0,\displaystyle\int_{\Gamma}\partial_{t}c\,v\,ds-\int_{\Gamma}c\mathbf{u}\cdot\nabla_{\Gamma}v\,ds+\int_{\Gamma}M\nabla_{\Gamma}\mu\cdot\nabla_{\Gamma}v\,ds=0, (2.16)
Γμg𝑑s=Γ1ϵf0(c)g𝑑s+ΓϵΓcΓgds,\displaystyle\int_{\Gamma}\mu\,g\,ds=\int_{\Gamma}\frac{1}{\epsilon}f_{0}^{\prime}(c)\,g\,ds+\int_{\Gamma}\epsilon\nabla_{\Gamma}c\cdot\nabla_{\Gamma}g\,ds, (2.17)

for all (𝐯,q,v,g)𝐕T×L2(Γ)×H1(Γ)×H1(Γ)(\mathbf{v},q,v,g)\in\mathbf{V}_{T}\times L^{2}(\Gamma)\times H^{1}(\Gamma)\times H^{1}(\Gamma).

Remark 2.1

By testing (2.14) with 𝐯=𝐮\mathbf{v}=\mathbf{u}, (2.15) with q=pq=p, (2.16) with v=μv=\mu, and using (2.7), one readily obtains the following energy equality:

ddtΓ(ρ2|𝐮|2+σγ(1ϵf0+ϵ2|Γc|2))𝑑s+Γ2η|Es(𝐮)|2𝑑s+ΓσγM|Γμ|2𝑑s=Γ𝐟𝐮𝑑s.\frac{d}{dt}\int_{\Gamma}\left(\frac{\rho}{2}|\mathbf{u}|^{2}+\sigma_{\gamma}\left(\frac{1}{\epsilon}f_{0}+\frac{\epsilon}{2}|\nabla_{\Gamma}c|^{2}\right)\right)ds+\int_{\Gamma}2\eta|E_{s}(\mathbf{u})|^{2}ds+\int_{\Gamma}\sigma_{\gamma}M|\nabla_{\Gamma}\mu|^{2}ds=\int_{\Gamma}\mathbf{f}\cdot\mathbf{u}\,ds. (2.18)

In other words, model (2.9), (2.5)–(2.7) is thermodynamically consistent, i.e. the system is dissipative for 𝐟=𝟎\mathbf{f}=\boldsymbol{0}.

Indeed, the first two terms in eq. (2.14) tested with 𝐯=𝐮\mathbf{v}=\mathbf{u} can be handled as follows:

Γ(ρt𝐮𝐮+ρ(Γ𝐮)𝐮𝐮)𝑑s=Γ(ρ2t|𝐮|212divΓ(ρ𝐮)|𝐮|2)𝑑s=Γ(12t(ρ|𝐮|2)12(tρ+divΓ(ρ𝐮))|𝐮|2)𝑑s.\begin{split}\int_{\Gamma}\left(\rho\partial_{t}\mathbf{u}\cdot\mathbf{u}+\rho(\nabla_{\Gamma}\mathbf{u})\mathbf{u}\cdot\mathbf{u}\right)ds&=\int_{\Gamma}\left(\frac{\rho}{2}\partial_{t}|\mathbf{u}|^{2}-\frac{1}{2}{\mathop{\,\rm div}}_{\Gamma}(\rho\mathbf{u})|\mathbf{u}|^{2}\right)ds\\ &=\int_{\Gamma}\left(\frac{1}{2}\partial_{t}(\rho|\mathbf{u}|^{2})-\frac{1}{2}\left(\partial_{t}\rho+{\mathop{\,\rm div}}_{\Gamma}(\rho\mathbf{u})\right)|\mathbf{u}|^{2}\right)ds.\end{split}

Using generic ρ=ρ(c)\rho=\rho(c), (2.8), and (2.5) we get

tρ+divΓ(ρ𝐮)=dρdc(tc+divΓ(c𝐮))=θ2(tc+divΓ(c𝐮)).\partial_{t}\rho+{\mathop{\,\rm div}}_{\Gamma}(\rho\mathbf{u})=\frac{d\rho}{dc}\left(\partial_{t}c+{\mathop{\,\rm div}}_{\Gamma}(c\mathbf{u})\right)=\theta^{2}\left(\partial_{t}c+{\mathop{\,\rm div}}_{\Gamma}(c\mathbf{u})\right).

With the help of (2.6) , this yields

Γ(tρ+divΓ(ρ𝐮))|𝐮|2𝑑s\displaystyle\int_{\Gamma}\left(\partial_{t}\rho+{\mathop{\,\rm div}}_{\Gamma}(\rho\mathbf{u})\right)|\mathbf{u}|^{2}ds =ΓM(Γμ)(Γ|θ𝐮|2)𝑑s\displaystyle=-\int_{\Gamma}M(\nabla_{\Gamma}\mu)\cdot(\nabla_{\Gamma}|\theta\mathbf{u}|^{2})ds
=2ΓM(Γ(θ𝐮))(Γμ)(θ𝐮)𝑑s,\displaystyle=-2\int_{\Gamma}M(\nabla_{\Gamma}(\theta\mathbf{u}))(\nabla_{\Gamma}\mu)\cdot(\theta\mathbf{u})ds,

which will cancel with the second term at the right-hand side in eq. (2.14) tested with 𝐯=𝐮\mathbf{v}=\mathbf{u}. Thus, from eq. (2.14) tested with 𝐯=𝐮\mathbf{v}=\mathbf{u} we obtain the following equality:

12ddtΓρ|𝐮|2𝑑s+Γ2η|Es(𝐮)|2𝑑s=ΓσγcΓμ𝐮ds+Γ𝐟𝐮𝑑s,\frac{1}{2}\frac{d}{dt}\int_{\Gamma}\rho|\mathbf{u}|^{2}ds+\int_{\Gamma}2\eta|E_{s}(\mathbf{u})|^{2}ds=-\int_{\Gamma}\sigma_{\gamma}c\nabla_{\Gamma}\mu\cdot\mathbf{u}\,ds+\int_{\Gamma}\mathbf{f}\cdot\mathbf{u}\,ds, (2.19)

where we have also used eq. (2.15) tested with q=pq=p. From eq. (2.16) tested with v=μv=\mu and multiplied by σγ\sigma_{\gamma}, we get:

Γσγ(𝐮Γc)μ𝑑s=Γσγ(𝐮Γμ)c𝑑s=ΓσγtcμdsΓσγM|Γμ|2𝑑s.\displaystyle\int_{\Gamma}\sigma_{\gamma}(\mathbf{u}\cdot\nabla_{\Gamma}c)\,\mu\,ds=-\int_{\Gamma}\sigma_{\gamma}(\mathbf{u}\cdot\nabla_{\Gamma}\mu)\,c\,ds=-\int_{\Gamma}\sigma_{\gamma}\partial_{t}c\,\mu\,ds-\int_{\Gamma}\sigma_{\gamma}M|\nabla_{\Gamma}\mu|^{2}ds. (2.20)

The first term on the right-hand side can be handled using (2.7) as follows:

Γtcμds=Γtc(1ϵf0ϵΔΓc)ds=ddtΓ(1ϵf0+ϵ2|Γc|2)𝑑s.\int_{\Gamma}\partial_{t}c\,\mu\,ds=\int_{\Gamma}\partial_{t}c\left(\frac{1}{\epsilon}f_{0}^{\prime}-\epsilon\Delta_{\Gamma}c\right)ds=\frac{d}{dt}\int_{\Gamma}\left(\frac{1}{\epsilon}f_{0}+\frac{\epsilon}{2}|\nabla_{\Gamma}c|^{2}\right)ds.

Plugging this into (2.20) and then using what we get in (2.19), we obtain the energy balance (2.18).

3 Numerical method

For the discretization of the variational problem (2.14)–(2.17) we apply an unfitted finite element method called trace finite element approach (TraceFEM). To discretize surface equations, TraceFEM relies on a tessellation of a 3D bulk computational domain Ω\Omega (ΓΩ\Gamma\subset\Omega holds) into shape regular tetrahedra untangled to the position of Γ\Gamma.

We start with required definitions to set up finite element spaces and variational form. A few auxiliary results will be proved. Then we proceed to the fully-discrete method by introducing a splitting time discretization, which decouples (2.14)–(2.17) into one linear fluid problem and one phase-field problem per every time step.

Surface Γ\Gamma is defined implicitly as the zero level set of a sufficiently smooth (at least Lipschitz continuous) function ϕ\phi, i.e. Γ={𝐱Ω:ϕ(𝐱)=0}\Gamma=\{\mathbf{x}\in\Omega\,:\,\phi(\mathbf{x})=0\}, such that |ϕ|c0>0|\nabla\phi|\geq c_{0}>0 in a 3D neighborhood U(Γ)U(\Gamma) of the surface. The vector field 𝐧=ϕ/|ϕ|\mathbf{n}=\nabla\phi/|\nabla\phi| is normal on Γ\Gamma and defines quasi-normal directions in U(Γ)U(\Gamma). Let 𝒯h\mathcal{T}_{h} be the collection of all tetrahedra such that Ω¯=T𝒯hT¯\overline{\Omega}=\cup_{T\in\mathcal{T}_{h}}\overline{T}. The subset of tetrahedra that have a nonzero intersection with Γ\Gamma is denoted by 𝒯hΓ\mathcal{T}_{h}^{\Gamma}. The grid can be refined towards Γ\Gamma, however the tetrahedra from 𝒯hΓ\mathcal{T}_{h}^{\Gamma} form a quasi-uniform tessellation with the characteristic tetrahedra size hh. The domain formed by all tetrahedra in 𝒯hΓ\mathcal{T}_{h}^{\Gamma} is denoted by ΩhΓ\Omega^{\Gamma}_{h}.

On 𝒯hΓ\mathcal{T}_{h}^{\Gamma} we use a standard finite element space of continuous functions that are piecewise-polynomials of degree kk. This bulk (volumetric) finite element space is denoted by VhkV_{h}^{k}:

Vhk={vC(ΩhΓ):vPk(T)for anyT𝒯hΓ},V_{h}^{k}=\{v\in C(\Omega^{\Gamma}_{h})\,:\,v\in P_{k}(T)~{}\text{for any}~{}T\in\mathcal{T}_{h}^{\Gamma}\},

In the trace finite element method formulated below, the traces of functions from VhkV_{h}^{k} on Γ\Gamma are used to approximate the surface fraction and the chemical potential. Our bulk velocity and pressure finite element spaces are either Taylor–Hood elements on ΩhΓ\Omega^{\Gamma}_{h},

𝐕h=(Vhm+1)3,Qh=VhmL02(Γ),\mathbf{V}_{h}=(V_{h}^{m+1})^{3},\quad Q_{h}=V_{h}^{m}\cap L^{2}_{0}(\Gamma), (3.1)

or equal order velocity–pressure elements

𝐕h=(Vhm)3,Qh=VhmL02(Γ),m1.\mathbf{V}_{h}=(V_{h}^{m})^{3},\quad Q_{h}=V_{h}^{m}\cap L^{2}_{0}(\Gamma),\quad m\geq 1. (3.2)

These spaces are employed to discretize the surface Navier–Stokes system. Surface velocity and pressure will be represented by the traces of functions from 𝐕h\mathbf{V}_{h} and QhQ_{h} on Γ\Gamma. In general, approximation orders kk (for the phase-field problem) and mm (for the fluid problem) can be chosen to be different.

Assumption 3.1

We assume that integrals over Γ\Gamma can be computed exactly, i.e. we do not consider geometry errors.

In practice, Γ\Gamma has to be approximated by a (sufficiently accurate) “discrete” surface ΓhΓ\Gamma_{h}\approx\Gamma in such a way that integrals over Γh\Gamma_{h} can be computed accurately and efficiently. For first order finite elements (m=1m=1, k=1k=1), a straightforward polygonal approximation of Γ\Gamma ensures that the geometric approximation error is consistent with the finite element interpolation error; see, e.g., [36]. For higher order elements, numerical approximation of Γ\Gamma based on, e.g., isoparametric trace FE can be used to recover the optimal accuracy [16] in a practical situation when parametrization of Γ\Gamma is not available explicitly.

There are two well-known issues related to the fact that we are dealing with surface and unfitted finite elements:

  1. 1.

    The numerical treatment of condition 𝐮𝐧\mathbf{u}\cdot\mathbf{n}. Enforcing the condition 𝐮h𝐧=0\mathbf{u}_{h}\cdot\mathbf{n}=0 on Γ\Gamma for polynomial functions 𝐮h𝐕h\mathbf{u}_{h}\in\mathbf{V}_{h} is inconvenient and may lead to locking (i.e., only 𝐮h=0\mathbf{u}_{h}=0 satisfies it). Instead, we add a penalty term to the weak formulation to enforce the tangential constraint weakly.

  2. 2.

    Possible small cuts of tetrahedra from 𝒯hΓ\mathcal{T}_{h}^{\Gamma} by the surface. For the standard choice of finite element basis functions, this may lead to poorly conditioned algebraic systems. We recover algebraic stability by adding certain volumetric terms to the finite element formulation.

To make the presentation of the finite element formulation more compact, we introduce the following finite element bilinear forms related to the Cahn–Hilliard part of the problem:

aμ(μ,v)ΓMΓμΓvds+τμΩhΓ(𝐧μ)(𝐧v)𝑑𝐱\displaystyle a_{\mu}(\mu,v)\coloneqq\int_{\Gamma}M\nabla_{\Gamma}\mu\cdot\nabla_{\Gamma}v\,ds+\tau_{\mu}\int_{\Omega^{\Gamma}_{h}}(\mathbf{n}\cdot\nabla\mu)(\mathbf{n}\cdot\nabla v)\,d\mathbf{x} (3.3)
ac(c,g)ϵΓΓcΓgds+τcΩhΓ(𝐧c)(𝐧g)𝑑𝐱.\displaystyle a_{c}(c,g)\coloneqq\epsilon\int_{\Gamma}\nabla_{\Gamma}c\cdot\nabla_{\Gamma}g\,ds+\tau_{c}\int_{\Omega^{\Gamma}_{h}}(\mathbf{n}\cdot\nabla c)(\mathbf{n}\cdot\nabla g)\,d\mathbf{x}. (3.4)

Forms (3.3)–(3.4) are well defined for μ,v,c,gH1(ΩhΓ)\mu,v,c,g\in H^{1}(\Omega^{\Gamma}_{h}). The second term in (3.3) and (3.4) take care of the issue of the small element cuts (the above issue 2) [16]. These terms are consistent up to geometric errors related to the approximation of Γ\Gamma by Γh\Gamma_{h} and 𝐧\mathbf{n} by 𝐧h\mathbf{n}_{h} in the following sense: any smooth solution μ\mu and cc can be extended off the surface along (quasi)-normal directions so that 𝐧μ=0\mathbf{n}\cdot\nabla\mu=0 and 𝐧c=0\mathbf{n}\cdot\nabla c=0 in ΩhΓ\Omega^{\Gamma}_{h}. We set the stabilization parameters as follows:

τμ=h,τc=ϵh1.\tau_{\mu}=h,\quad\tau_{c}=\epsilon\,h^{-1}.

For the stability of a numerical method, it is crucial that the computed density and viscosity stay positive. The polynomial double-well potential does not enforce cc to stay within [0,1][0,1] interval and hence ρ\rho and η\eta may eventually take negative values, if one adopts the linear relation between cc and ρ\rho in (2.1) or between cc and η\eta in (2.2). Numerical errors may be another reason for the order parameter to depart significantly from [0,1][0,1]. Therefore, assuming without loss of generality that ρ1ρ2\rho_{1}\geq\rho_{2} and η1η2\eta_{1}\geq\eta_{2}, we first replace (2.1) and (2.2) with the following cut-off functions that ensure density and viscosity stay positive:

ρ(c)={ρ2c0cρ1+(1c)ρ2c>0η(c)={η2c0cη1+(1c)η2c>0\rho(c)=\left\{\begin{array}[]{cc}\rho_{2}&c\leq 0\\ c\rho_{1}+(1-c)\rho_{2}&c>0\end{array}\right.\qquad\eta(c)=\left\{\begin{array}[]{cc}\eta_{2}&c\leq 0\\ c\eta_{1}+(1-c)\eta_{2}&c>0\end{array}\right. (3.5)

Note that unlike some previous studies, we clip the linear dependence (2.1) and (2.2) only from below. The resulting convexity of ρ(c)\rho(c) plays a role in the stability analysis later. Nevertheless, (3.5) is not completely satisfactory since θ2=δρδc\theta^{2}=\frac{\delta\rho}{\delta c} is discontinuous, while we need θ\theta to be from C1C^{1}. To this end, we approximate ρ(c)\rho(c) from (2.1) by a smooth monotone convex and uniformly positive function. In our implementation we let θ2=ρ1ρ22(tanh(c/α)+1)\theta^{2}=\frac{\rho_{1}-\rho_{2}}{2}\left(\tanh(c/\alpha)+1\right), with α=0.1\alpha=0.1, and ρ(c)=0cθ2(t)𝑑t+ρ2\rho(c)=\int_{0}^{c}\theta^{2}(t)dt+\rho_{2}.

Later we make use of the decomposition of a vector field on Γ\Gamma into its tangential and normal components: 𝐮=𝐮¯+(𝐮𝐧)𝐧.\mathbf{u}=\overline{\mathbf{u}}+(\mathbf{u}\cdot\mathbf{n})\mathbf{n}. For the Navier–Stokes part, we introduce the following forms:

a(η;𝐮,𝐯)Γ2ηEs(𝐮¯):Es(𝐯¯)ds+τΓ(𝐧𝐮)(𝐧𝐯)𝑑s+βuΩhΓ[(𝐧)𝐮][(𝐧)𝐯]𝑑𝐱,\displaystyle a(\eta;\mathbf{u},\mathbf{v})\coloneqq\int_{\Gamma}2\eta E_{s}(\overline{\mathbf{u}}):E_{s}(\overline{\mathbf{v}})\,ds+\tau\int_{\Gamma}(\mathbf{n}\cdot\mathbf{u})(\mathbf{n}\cdot\mathbf{v})\,ds+\beta_{u}\int_{\Omega^{\Gamma}_{h}}[(\mathbf{n}\cdot\nabla)\mathbf{u}]\cdot[(\mathbf{n}\cdot\nabla)\mathbf{v}]\,d\mathbf{x}, (3.6)
c(ρ;𝐰,𝐮,𝐯)Γρ𝐯T(Γ𝐮¯)𝐰𝑑s+12Γρ^(divΓ𝐰¯)𝐮¯𝐯¯𝑑s,\displaystyle c(\rho;\mathbf{w},\mathbf{u},\mathbf{v})\coloneqq\int_{\Gamma}\rho\mathbf{v}^{T}(\nabla_{\Gamma}\overline{\mathbf{u}})\mathbf{w}\,ds+\frac{1}{2}\int_{\Gamma}\widehat{\rho}({\mathop{\,\rm div}}_{\Gamma}\overline{\mathbf{w}})\overline{\mathbf{u}}\cdot\overline{\mathbf{v}}\,ds, (3.7)
b(𝐮,q)=Γ𝐮Γqds,\displaystyle b(\mathbf{u},q)=\int_{\Gamma}\mathbf{u}\cdot\nabla_{\Gamma}q\,ds, (3.8)
s(p,q)βpΩhΓ(𝐧p)(𝐧q)𝑑𝐱.\displaystyle s(p,q)\coloneqq\beta_{p}\int_{\Omega^{\Gamma}_{h}}(\mathbf{n}\cdot\nabla p)(\mathbf{n}\cdot\nabla q)\,d\mathbf{x}. (3.9)

with ρ^=ρdρdcc\widehat{\rho}=\rho-\frac{d\rho}{d\,c}c. Forms (3.6)–(3.9) are well defined for p,qH1(ΩhΓ)p,q\in H^{1}(\Omega^{\Gamma}_{h}), 𝐮,𝐯,𝐰H1(ΩhΓ)3\mathbf{u},\mathbf{v},\mathbf{w}\in H^{1}(\Omega^{\Gamma}_{h})^{3}. In (3.6), τ>0\tau>0 is a penalty parameter to enforce the tangential constraint (i.e., to address the above issue 1), while βu0\beta_{u}\geq 0 in (3.6) and βp0\beta_{p}\geq 0 in (3.9) are stabilization parameters to deal with possible small cuts. They are set according to [24]:

τ=h2,βp=h,βu=h1.\tau=h^{-2},\quad\beta_{p}=h,\quad\beta_{u}=h^{-1}. (3.10)

If one uses equal order pressure-velocity trace elements instead of Taylor–Hood elements, then form (3.9) should be replaced by

s(p,q)βpΩhΓpqd𝐱.\displaystyle s(p,q)\coloneqq\beta_{p}\int_{\Omega^{\Gamma}_{h}}\nabla p\cdot\nabla q\,d\mathbf{x}.

The second term in (3.7) is consistent since the divergence of the true tangential velocity is zero. To avoid differentiation of the projector operator, one may use the identity Γ𝐮¯=Γ𝐮(𝐮𝐧)𝐇\nabla_{\Gamma}\overline{\mathbf{u}}=\nabla_{\Gamma}\mathbf{u}-(\mathbf{u}\cdot\mathbf{n})\mathbf{H} to implement the aa-form and cc-form. For the analysis, we need the identity for the form (3.7) given in the following elementary lemma.

Lemma 3.1

For any 𝐮,𝐰H1(ΩhΓ)3\mathbf{u},\mathbf{w}\in H^{1}(\Omega^{\Gamma}_{h})^{3}, it holds

c(ρ;𝐰,𝐮,𝐮)=12Γθ2divΓ(c𝐰¯)|𝐮¯|2ds.c(\rho;\mathbf{w},\mathbf{u},\mathbf{u})=-\frac{1}{2}\int_{\Gamma}\theta^{2}{\mathop{\,\rm div}}_{\Gamma}(c\overline{\mathbf{w}})|\overline{\mathbf{u}}|^{2}\,ds.

Proof:

Using the definition of the covariant gradient in terms of tangential operators, Γ𝐮¯=𝐏(𝐮¯)𝐏\nabla_{\Gamma}\overline{\mathbf{u}}=\mathbf{P}(\nabla\overline{\mathbf{u}})\mathbf{P}, and the integration by parts (2.13), we compute for the first integral term in (3.7)

Γρ𝐮T(Γ𝐮¯)𝐰𝑑s=Γρ𝐮¯T(Γ𝐮¯)𝐰¯𝑑s=ΓdivΓ(ρ𝐮¯𝐰¯)𝐮¯ds=ΓρdivΓ(𝐰¯)|𝐮¯|2dsΓ𝐮¯Γ(ρ𝐮¯)𝐰¯ds=ΓρdivΓ(𝐰¯)|𝐮¯|2dsΓρ𝐮T(Γ𝐮¯)𝐰𝑑sΓ(𝐰¯Γρ)|𝐮¯|2𝑑s\begin{split}\int_{\Gamma}\rho\mathbf{u}^{T}(\nabla_{\Gamma}\overline{\mathbf{u}})\mathbf{w}\,ds&=\int_{\Gamma}\rho\overline{\mathbf{u}}^{T}(\nabla_{\Gamma}\overline{\mathbf{u}})\overline{\mathbf{w}}\,ds=-\int_{\Gamma}{\mathop{\,\rm div}}_{\Gamma}(\rho\overline{\mathbf{u}}\otimes\overline{\mathbf{w}})\cdot\overline{\mathbf{u}}\,ds\\ &=-\int_{\Gamma}\rho{\mathop{\,\rm div}}_{\Gamma}(\overline{\mathbf{w}})|\overline{\mathbf{u}}|^{2}\,ds-\int_{\Gamma}\overline{\mathbf{u}}\nabla_{\Gamma}(\rho\overline{\mathbf{u}})\overline{\mathbf{w}}\,ds\\ &=-\int_{\Gamma}\rho{\mathop{\,\rm div}}_{\Gamma}(\overline{\mathbf{w}})|\overline{\mathbf{u}}|^{2}\,ds-\int_{\Gamma}\rho\mathbf{u}^{T}(\nabla_{\Gamma}\overline{\mathbf{u}})\mathbf{w}\,ds-\int_{\Gamma}(\overline{\mathbf{w}}\cdot\nabla_{\Gamma}\rho)|\overline{\mathbf{u}}|^{2}\,ds\end{split}

From this equality and using Γρ=dρdcΓc\nabla_{\Gamma}\rho=\frac{d\rho}{dc}\nabla_{\Gamma}c we obtain:

Γρ𝐮T(Γ𝐮¯)𝐰𝑑s=12ΓρdivΓ(𝐰¯)|𝐮¯|2ds12Γdρdc(𝐰¯Γc)|𝐮¯|2𝑑s\int_{\Gamma}\rho\mathbf{u}^{T}(\nabla_{\Gamma}\overline{\mathbf{u}})\mathbf{w}\,ds=-\frac{1}{2}\int_{\Gamma}\rho{\mathop{\,\rm div}}_{\Gamma}(\overline{\mathbf{w}})|\overline{\mathbf{u}}|^{2}\,ds-\frac{1}{2}\int_{\Gamma}\frac{d\rho}{d\,c}(\overline{\mathbf{w}}\cdot\nabla_{\Gamma}c)|\overline{\mathbf{u}}|^{2}\,ds (3.11)

We recall that ρ^=ρdρdcc\widehat{\rho}=\rho-\frac{d\rho}{d\,c}c and substitute (3.11) into the definition of the form (3.7). After straightforward computations, we arrive at the result:

c(ρ;𝐰,𝐮,𝐮)=Γρ𝐮T(Γ𝐮¯)𝐰ds+12Γ(ρdρdcc)(divΓ𝐰¯)|𝐮¯|2ds=12ΓρdivΓ(𝐰¯)|𝐮¯|2ds12Γdρdc(𝐰¯Γc)|𝐮¯|2𝑑s+12Γ(ρdρdcc)(divΓ𝐰¯)|𝐮¯|2𝑑s=12Γdρdc(𝐰¯Γc)|𝐮¯|2𝑑s12Γdρdcc(divΓ𝐰¯)|𝐮¯|2𝑑s=12ΓdρdcdivΓ(c𝐰¯)|𝐮¯|2ds=12Γθ2divΓ(c𝐰¯)|𝐮¯|2ds.\begin{split}c(\rho;\mathbf{w},&\mathbf{u},\mathbf{u})=\int_{\Gamma}\rho\mathbf{u}^{T}(\nabla_{\Gamma}\overline{\mathbf{u}})\mathbf{w}\,ds+\frac{1}{2}\int_{\Gamma}(\rho-\frac{d\rho}{d\,c}c)({\mathop{\,\rm div}}_{\Gamma}\overline{\mathbf{w}})|\overline{\mathbf{u}}|^{2}\,ds\\ &=-\frac{1}{2}\int_{\Gamma}\rho{\mathop{\,\rm div}}_{\Gamma}(\overline{\mathbf{w}})|\overline{\mathbf{u}}|^{2}\,ds-\frac{1}{2}\int_{\Gamma}\frac{d\rho}{d\,c}(\overline{\mathbf{w}}\cdot\nabla_{\Gamma}c)|\overline{\mathbf{u}}|^{2}\,ds+\frac{1}{2}\int_{\Gamma}(\rho-\frac{d\rho}{d\,c}c)({\mathop{\,\rm div}}_{\Gamma}\overline{\mathbf{w}})|\overline{\mathbf{u}}|^{2}\,ds\\ &=-\frac{1}{2}\int_{\Gamma}\frac{d\rho}{d\,c}(\overline{\mathbf{w}}\cdot\nabla_{\Gamma}c)|\overline{\mathbf{u}}|^{2}\,ds-\frac{1}{2}\int_{\Gamma}\frac{d\rho}{d\,c}c({\mathop{\,\rm div}}_{\Gamma}\overline{\mathbf{w}})|\overline{\mathbf{u}}|^{2}\,ds\\ &=-\frac{1}{2}\int_{\Gamma}\frac{d\rho}{d\,c}{\mathop{\,\rm div}}_{\Gamma}(c\overline{\mathbf{w}})|\overline{\mathbf{u}}|^{2}\,ds=-\frac{1}{2}\int_{\Gamma}\theta^{2}{\mathop{\,\rm div}}_{\Gamma}(c\overline{\mathbf{w}})|\overline{\mathbf{u}}|^{2}\,ds.\end{split}

\square

For the numerical experiments in Sec. 5, we also add to bilinear form (3.6) the grad-div stabilization term [35], γ^ΓdivΓ𝐮divΓ𝐯ds\widehat{\gamma}\,\int_{\Gamma}{\mathop{\,\rm div}}_{\Gamma}\mathbf{u}\,{\mathop{\,\rm div}}_{\Gamma}\mathbf{v}\,ds. This term is not essential, in particular for the analysis in Sec. 4, but we find it beneficial for the performance of the iterative algebraic solver and for the overall accuracy of the solution. We set the grad-div stabilization parameter γ^=1\widehat{\gamma}=1.

At time instance tn=nΔtt^{n}=n\Delta t, with time step Δt=TN\Delta t=\frac{T}{N}, ζn\zeta^{n} denotes the approximation of generic variable ζ(tn,𝐱)\zeta(t^{n},\mathbf{x}). Further, we introduce the following notation for a first order approximation of the time derivative:

[ζ]tn=ζnζn1Δt.\left[\zeta\right]_{t}^{n}=\frac{\zeta^{n}-\zeta^{n-1}}{\Delta t}. (3.12)

The decoupled linear finite element method analyzed and tested in this paper reads as follows. At time step tn+1t^{n+1}, perform:

  • -

    Step 1: Given 𝐮hn𝐕h\mathbf{u}^{n}_{h}\in\mathbf{V}_{h} and chnVhkc^{n}_{h}\in V_{h}^{k}, find (chn+1,μhn+1)Vhk×Vhk(c^{n+1}_{h},\mu^{n+1}_{h})\in V^{k}_{h}\times V^{k}_{h} such that:

    ([ch]tn+1,vh)(𝐮hnchn+1,Γvh)+aμ(μhn+1,vh)=0,\displaystyle\left(\left[c_{h}\right]_{t}^{n+1},v_{h}\right)-\left(\mathbf{u}^{n}_{h}c^{n+1}_{h},\nabla_{\Gamma}v_{h}\right)+a_{\mu}(\mu_{h}^{n+1},v_{h})=0, (3.13)
    (μhn+1γcΔtϵ[ch]tn+11ϵf0(chn),gh)ac(chn+1,gh)=0,\displaystyle\left(\mu_{h}^{n+1}-\frac{\gamma_{c}\Delta t}{\epsilon}\left[c_{h}\right]_{t}^{n+1}-\frac{1}{\epsilon}f^{\prime}_{0}(c_{h}^{n}),\,g_{h}\right)-a_{c}(c_{h}^{n+1},g_{h})=0, (3.14)

    for all (vh,gh)Vhk×Vhk(v_{h},g_{h})\in V^{k}_{h}\times V^{k}_{h}.

  • -

    Step 2: Set θn+1=dρdc(chn+1)\theta^{n+1}=\sqrt{\frac{d\rho}{dc}(c^{n+1}_{h})}. Find (𝐮hn+1,phn+1)𝐕h×Qh(\mathbf{u}_{h}^{n+1},p_{h}^{n+1})\in\mathbf{V}_{h}\times Q_{h} such that

    (ρn[𝐮¯h]tn+1,𝐯h)+c(ρn+1;𝐮hn,𝐮hn+1,𝐯h)+a(ηn+1;𝐮hn+1,𝐯h)+b(𝐯h,phn+1)\displaystyle(\rho^{n}\left[\overline{\mathbf{u}}_{h}\right]_{t}^{n+1},\mathbf{v}_{h})+c(\rho^{n+1};\mathbf{u}^{n}_{h},\mathbf{u}^{n+1}_{h},\mathbf{v}_{h})+a(\eta^{n+1};\mathbf{u}_{h}^{n+1},\mathbf{v}_{h})+b(\mathbf{v}_{h},p_{h}^{n+1})
    =(σγchn+1Γμhn+1,𝐯h)+M((Γ(θn+1𝐮¯hn+1))Γμhn+1,θn+1𝐯h)+(𝐟hn+1,𝐯h)\displaystyle\quad\quad=-(\sigma_{\gamma}c^{n+1}_{h}\nabla_{\Gamma}\mu^{n+1}_{h},\mathbf{v}_{h})+M\left((\nabla_{\Gamma}(\theta^{n+1}\overline{\mathbf{u}}_{h}^{n+1}))\nabla_{\Gamma}\mu^{n+1}_{h},\theta^{n+1}\mathbf{v}_{h}\right)+(\mathbf{f}_{h}^{n+1},\mathbf{v}_{h}) (3.15)
    b(𝐮hn+1,qh)s(phn+1,qh)=0\displaystyle b(\mathbf{u}_{h}^{n+1},q_{h})-s(p_{h}^{n+1},q_{h})=0 (3.16)

    for all (𝐯h,qh)𝐕h×Qh(\mathbf{v}_{h},q_{h})\in\mathbf{V}_{h}\times Q_{h}.

At each sub-step of the scheme, a linear problem (Chan–Hilliard type system at step 1 and linearized Navier–Stokes system at step 2) has to be solved. This allows us to achieve low computational costs, while the scheme is provably stable under relatively mild restrictions. Moreover, the results of numerical experiments in Sec. 5 do not show that any restrictions on the discretization parameters are required in practice (Remark 4.3 discusses what arguments in our analysis require these restrictions).

Before proceeding with analysis, we note that the above scheme does not modify the advection velocity in (3.13) for the purpose of analysis, unlike some other linear decoupled schemes for the NSCH equations found in the literature. We also avoid another common helpful trick to prove energy stability of the variable density NSCH, namely the modification of the momentum equation based on a mass conservation condition of the form tρ+𝐮ΓcdivΓ(MΓμ)=0,\partial_{t}\rho+\mathbf{u}\cdot\nabla_{\Gamma}c-{\mathop{\,\rm div}}_{\Gamma}\left(M\nabla_{\Gamma}\mu\right)=0, which follow from (2.6) and (2.1). This modification, however, is not consistent if a non-linear relation between ρ\rho and cc is adopted; also it introduces several extra terms in the finite element formulation.

4 Analysis of the decoupled Finite Element method

For the analysis in this section, we assume no forcing term, i.e. 𝐟n+1=𝟎\mathbf{f}^{n+1}=\boldsymbol{0} for all nn. To avoid technical complications related to handling Killing vector fields (see, e.g. discussion in [5]), we shall also assume that the surface does not support any tangential rigid motions, and so 𝐕T0=𝐕T\mathbf{V}_{T}^{0}=\mathbf{V}_{T}.

We further use the aba\lesssim b notation if inequality acba\leq c\,b holds between quantities aa and bb with a constant cc independent of hh, Δt\Delta t, ϵ\epsilon, and the position of Γ\Gamma in the background mesh. We give similar meaning to aba\gtrsim b, and aba\simeq b means that both aba\lesssim b and aba\gtrsim b hold.

The following lemma will be useful in the proof of the main result, which is reported in Theorem 4.2.

Lemma 4.1

It holds

vhL(Γ)|lnh|12vhH1(Γ)+h12𝐧vhL2(ΩhΓ)vhVhk.\|v_{h}\|_{L^{\infty}(\Gamma)}\lesssim|\ln h|^{\frac{1}{2}}\|v_{h}\|_{H^{1}(\Gamma)}+h^{-\frac{1}{2}}\left\|\mathbf{n}\cdot\nabla v_{h}\right\|_{L^{2}(\Omega_{h}^{\Gamma})}\qquad\forall\,v_{h}\in V_{h}^{k}. (4.1)

For vhv_{h} satisfying Γvh𝑑s=0\int_{\Gamma}v_{h}\,ds=0, the factor vhH1(Γ)\|v_{h}\|_{H^{1}(\Gamma)} on the r.h.s. can be replaced by ΓvhL2(Γ)\|\nabla_{\Gamma}v_{h}\|_{L^{2}(\Gamma)}.

Proof:

Denote by 𝐩(𝐱)Γ\mathbf{p}(\mathbf{x})\in\Gamma, 𝐱Ω\mathbf{x}\in\Omega, the closest point projection on Γ\Gamma. Since Γ\Gamma is smooth, there is a tubular neighborhood of Γ\Gamma

U={𝐱3:dist(𝐱,Γ)<d},d>0,U=\{\mathbf{x}\in\mathbb{R}^{3}\,:\,\mbox{dist}(\mathbf{x},\Gamma)<d\},\quad d>0,

such that 𝐱=𝐩(𝐱)+s𝐧\mathbf{x}=\mathbf{p}(\mathbf{x})+s\mathbf{n}, 𝐱U\mathbf{x}\in U, defines the local coordinate system (s,𝐲)(s,\mathbf{y}), with 𝐲=𝐩(𝐱)\mathbf{y}=\mathbf{p}(\mathbf{x}) and |s|=dist(𝐱,Γ)|s|=\mbox{dist}(\mathbf{x},\Gamma). We can always assume hh0=O(1)h\leq h_{0}=O(1), such that ΩhΓU\Omega_{h}^{\Gamma}\subset U. For uH1(U)u\in H^{1}(U), we have in UU

u(𝐱)=u(𝐲)+0s𝐧u(r,𝐲)𝑑r.u(\mathbf{x})=u(\mathbf{y})+\int_{0}^{s}\mathbf{n}\cdot\nabla u(r,\mathbf{y})\,dr. (4.2)

Denote by Ω~hΓ\widetilde{\Omega}_{h}^{\Gamma} a “reachable from Γ\Gamma” part of ΩhΓ\Omega_{h}^{\Gamma} in the following sense: for any 𝐱Ω~hΓ\mathbf{x}\in\widetilde{\Omega}_{h}^{\Gamma} the interval (𝐱,𝐩(𝐱))(\mathbf{x},\mathbf{p}(\mathbf{x})) is completely inside ΩhΓ\Omega_{h}^{\Gamma}. Let

g±(𝐲)=±sup{s:(𝐲±s𝐧,𝐲)ΩhΓ},𝐲Γ,g_{\pm}(\mathbf{y})=\pm\sup\{s\in\mathbb{R}\,:\,(\mathbf{y}\pm s\mathbf{n},\mathbf{y})\subset\Omega_{h}^{\Gamma}\},\quad\mathbf{y}\in\Gamma,

where g±(𝐲)g_{\pm}(\mathbf{y}) are piecewise smooth and g±L(Γ)h\|g_{\pm}\|_{L^{\infty}(\Gamma)}\lesssim h. Thanks to the co-area formula and the smoothness of Γ\Gamma, it holds

Ω~hΓ|f|𝑑𝐱Γgg+|f|𝑑𝐲𝑑s,for anyfL1(Ω~hΓ).\int_{\widetilde{\Omega}_{h}^{\Gamma}}|f|\,d\mathbf{x}\simeq\int_{\Gamma}\int_{g_{-}}^{g_{+}}|f|\,d\mathbf{y}\,ds,\quad\mbox{for any}~{}f\in L^{1}(\widetilde{\Omega}_{h}^{\Gamma}). (4.3)

From (4.2) and (4.3), we have:

u(𝐱)Lp(Ω~hΓ)(Γgg+|u(𝐲)+0s𝐧u(r,𝐲)𝑑r|p𝑑s𝑑𝐲)1p,\|u(\mathbf{x})\|_{L^{p}(\widetilde{\Omega}_{h}^{\Gamma})}\simeq\left(\int_{\Gamma}\int_{g_{-}}^{g_{+}}\left|u(\mathbf{y})+\int_{0}^{s}\mathbf{n}\cdot\nabla u(r,\mathbf{y})\,dr\right|^{p}\,ds\,d\mathbf{y}\right)^{\frac{1}{p}},

for any real exponent p>1p>1. Triangle inequality and inequality:

(Γgg+|u(𝐲)|p𝑑s𝑑𝐲)1p|g+g|1puLp(Γ)\left(\int_{\Gamma}\int_{g_{-}}^{g_{+}}\left|u(\mathbf{y})\right|^{p}\,ds\,d\mathbf{y}\right)^{\frac{1}{p}}\leq|g_{+}-g_{-}|^{\frac{1}{p}}\|u\|_{L^{p}(\Gamma)}

yield

uLp(Ω~hΓ)h1puLp(Γ)+(Γgg+|0s𝐧u(r,𝐲)𝑑r|p𝑑s𝑑𝐲)1p.\|u\|_{L^{p}(\widetilde{\Omega}_{h}^{\Gamma})}\lesssim h^{\frac{1}{p}}\|u\|_{L^{p}(\Gamma)}+\left(\int_{\Gamma}\int_{g_{-}}^{g_{+}}\left|\int_{0}^{s}\mathbf{n}\cdot\nabla u(r,\mathbf{y})\,dr\right|^{p}\,ds\,d\mathbf{y}\right)^{\frac{1}{p}}. (4.4)

To handle the second term on the right-hand side, we apply Hölder’s inequality with p=p/(p1)p^{\prime}=p/(p-1):

(Γgg+|0s𝐧u(r,𝐲)𝑑r|p𝑑s𝑑𝐲)1p(Γgg+|(0s|𝐧u(r,𝐲)|p𝑑r)1p(0s𝑑r)1p|p𝑑s𝑑𝐲)1p=(Γgg+(0s|𝐧u(r,𝐲)|p𝑑r)sp1𝑑s𝑑𝐲)1p(Γgg+|𝐧u(r,𝐲)|p𝑑r𝑑𝐲)1p(gg+sp1𝑑s)1ph𝐧uLp(Ω~hΓ)\begin{split}\left(\int_{\Gamma}\int_{g_{-}}^{g_{+}}\left|\int_{0}^{s}\mathbf{n}\cdot\nabla u(r,\mathbf{y})\,dr\right|^{p}\,ds\,d\mathbf{y}\right)^{\frac{1}{p}}&\leq\left(\int_{\Gamma}\int_{g_{-}}^{g_{+}}\left|\left(\int_{0}^{s}\left|\mathbf{n}\cdot\nabla u(r,\mathbf{y})\right|^{p}\,dr\right)^{\frac{1}{p}}\left(\int_{0}^{s}\,dr\right)^{\frac{1}{p^{\prime}}}\right|^{p}\,ds\,d\mathbf{y}\right)^{\frac{1}{p}}\\ &=\left(\int_{\Gamma}\int_{g_{-}}^{g_{+}}\left(\int_{0}^{s}\left|\mathbf{n}\cdot\nabla u(r,\mathbf{y})\right|^{p}\,dr\right)s^{p-1}\,ds\,d\mathbf{y}\right)^{\frac{1}{p}}\\ &\leq\left(\int_{\Gamma}\int_{g_{-}}^{g_{+}}\left|\mathbf{n}\cdot\nabla u(r,\mathbf{y})\right|^{p}\,dr\,d\mathbf{y}\right)^{\frac{1}{p}}\left(\int_{g_{-}}^{g_{+}}s^{p-1}\,ds\right)^{\frac{1}{p}}\\ &\lesssim h\left\|\mathbf{n}\cdot\nabla u\right\|_{L^{p}(\widetilde{\Omega}_{h}^{\Gamma})}\end{split}

Substituting this in (4.4) and using Ω~hΓΩhΓ\widetilde{\Omega}_{h}^{\Gamma}\subset\Omega_{h}^{\Gamma} we get

uLp(Ω~hΓ)h1pu(𝐱)Lp(Γ)+h𝐧uLp(ΩhΓ).\|u\|_{L^{p}(\widetilde{\Omega}_{h}^{\Gamma})}\lesssim h^{\frac{1}{p}}\|u(\mathbf{x})\|_{L^{p}(\Gamma)}+h\left\|\mathbf{n}\cdot\nabla u\right\|_{L^{p}(\Omega_{h}^{\Gamma})}.

Letting u=uhVhku=u_{h}\in V_{h}^{k} and applying FE inverse inequality to treat the last term, we arrive at

uhLp(Ω~hΓ))h1puhLp(Γ)+h3p12𝐧uhL2(ΩhΓ).\|u_{h}\|_{L^{p}(\widetilde{\Omega}_{h}^{\Gamma}))}\lesssim h^{\frac{1}{p}}\|u_{h}\|_{L^{p}(\Gamma)}+h^{\frac{3}{p}-\frac{1}{2}}\left\|\mathbf{n}\cdot\nabla u_{h}\right\|_{L^{2}(\Omega_{h}^{\Gamma})}. (4.5)

Next, we need the following technical result from Lemma 7.9 in [16]: There is δh\delta\simeq h, depending only on the shape regularity of the mesh 𝒯hΓ\mathcal{T}_{h}^{\Gamma} such that for any T𝒯hΓT\in\mathcal{T}_{h}^{\Gamma} there exists a ball Bδ(T)TΩ~hΓB_{\delta}(T)\subset T\cap\widetilde{\Omega}_{h}^{\Gamma} of radius δ\delta. Since on every tetrahedron uhu_{h} is polynomial function of a fixed degree by the standard norm equivalence argument, we have

(T|uh|p𝑑𝐱)1pC(Bδ(T)|uh|p𝑑𝐱)1p,\left(\int_{T}|u_{h}|^{p}\,d\mathbf{x}\right)^{\frac{1}{p}}\leq C\left(\int_{B_{\delta}(T)}|u_{h}|^{p}\,d\mathbf{x}\right)^{\frac{1}{p}},

with CC depending only on δ\delta and the shape regularity of TT. Raising both parts of this inequality to power pp, summing over all T𝒯hΓT\in\mathcal{T}_{h}^{\Gamma}, raising both parts to power 1/p1/p and using Bδ(T)Ω~hΓB_{\delta}(T)\subset\widetilde{\Omega}_{h}^{\Gamma} we get

uhLp(ΩhΓ)uhLp(Ω~hΓ)\|u_{h}\|_{L^{p}(\Omega_{h}^{\Gamma})}\lesssim\|u_{h}\|_{L^{p}(\widetilde{\Omega}_{h}^{\Gamma})}

We now use this in (4.5) and apply another FE inverse inequality to get

uL(Γ)uL(ΩhΓ)h3puhLp(ΩhΓ)h2puhLp(Γ)+h12𝐧uhL2(ΩhΓ).\|u\|_{L^{\infty}(\Gamma)}\leq\|u\|_{L^{\infty}(\Omega_{h}^{\Gamma})}\lesssim h^{-\frac{3}{p}}\|u_{h}\|_{L^{p}(\Omega_{h}^{\Gamma})}\lesssim h^{-\frac{2}{p}}\|u_{h}\|_{L^{p}(\Gamma)}+h^{-\frac{1}{2}}\left\|\mathbf{n}\cdot\nabla u_{h}\right\|_{L^{2}(\Omega_{h}^{\Gamma})}. (4.6)

For uH1(Γ)u\in H^{1}(\Gamma), recall that the Sobolev embedding theorem implies uLp(Γ)u\in L^{p}(\Gamma), p[1,)p\in[1,\infty), and

uLp(Γ)cp12uH1(Γ).\|u\|_{L^{p}(\Gamma)}\leq c\,p^{\frac{1}{2}}\|u\|_{H^{1}(\Gamma)}.

This result follows from the corresponding embedding theorem in 2\mathbb{R}^{2} by standard arguments based on the surface local parametrization and partition of unity. Using this in (4.6), we obtain the estimate

uL(Γ)h2pp12uhH1(Γ)+h12𝐧uhL2(ΩhΓ).\|u\|_{L^{\infty}(\Gamma)}\lesssim h^{-\frac{2}{p}}p^{\frac{1}{2}}\|u_{h}\|_{H^{1}(\Gamma)}+h^{-\frac{1}{2}}\left\|\mathbf{n}\cdot\nabla u_{h}\right\|_{L^{2}(\Omega_{h}^{\Gamma})}. (4.7)

Letting p=|lnh|p=|\ln h| proves the result in (4.1). Applying the Poincaré inequality for the function uhu_{h} satisfying Γuh𝑑s=0\int_{\Gamma}u_{h}\,ds=0 proves the second claim of the lemma. \square

Following [39], we modify the double-well potential in (2.10) for the purpose of analysis so that it is C2C^{2} smooth but has quadratic growth for large |c||c|. Straightforward calculations give us the following expression for f0(c)f_{0}^{\prime}(c) with sufficiently large but fixed α>1\alpha>1:

f0(c)={3α214c(α34+38α218),c>1+α2,(c232c+12)c,c[1α2,1+α2],3α214c+(α3438α2+18),c<1α2.\displaystyle{f}_{0}^{\prime}(c)=\begin{cases}\frac{3\alpha^{2}-1}{4}c-\left(\frac{\alpha^{3}}{4}+\frac{3}{8}\alpha^{2}-\frac{1}{8}\right),\quad c>\frac{1+\alpha}{2},\\ \left(c^{2}-\frac{3}{2}c+\frac{1}{2}\right)c,\quad c\in\left[\frac{1-\alpha}{2},\frac{1+\alpha}{2}\right],\\ \frac{3\alpha^{2}-1}{4}c+\left(\frac{\alpha^{3}}{4}-\frac{3}{8}\alpha^{2}+\frac{1}{8}\right),\quad c<\frac{1-\alpha}{2}.\end{cases}

Function f0(x)f_{0}^{\prime}(x) satisfies the following Lipschitz condition with L=3α214L=\frac{3\alpha^{2}-1}{4}:

14f0(x)f0(y)xyL,x,y,xy,\displaystyle-\frac{1}{4}\leq\frac{{f}_{0}^{\prime}(x)-{f}_{0}^{\prime}(y)}{x-y}\leq{}L,\quad\forall x,y\in\mathbb{R},~{}x\neq y, (4.8)

and growth condition

|f0(x)|L|x|.\displaystyle|{f}_{0}^{\prime}(x)|\leq L|x|. (4.9)
Theorem 4.2

Assume hh and Δt\Delta t satisfy Δtc|lnh|1ϵ\Delta t\leq c|\ln h|^{-1}\epsilon and hc|lnh|1min{Δt,|lnh|12ϵ|Δt|12}h\leq c|\ln h|^{-1}\min\{\Delta t,|\ln h|^{-\frac{1}{2}}\epsilon|\Delta t|^{\frac{1}{2}}\} for some sufficiently small c>0c>0, independent of hh, Δt\Delta t, ϵ\epsilon and position of Γ\Gamma in the background mesh. Then, it holds

Γ(ρN|𝐮¯hN|2+σγϵf0(chN))𝑑s+ac(chN,chN)+n=1NΔt(a(ηn;𝐮hn,𝐮hn)+aμ(μn,μn)+sh(phn,phn))K,\int_{\Gamma}\left(\rho^{N}|\overline{\mathbf{u}}_{h}^{N}|^{2}+\frac{\sigma_{\gamma}}{\epsilon}f_{0}(c^{N}_{h})\right)ds+a_{c}(c^{N}_{h},c^{N}_{h})+\sum_{n=1}^{N}\Delta t\left(a(\eta^{n};\mathbf{u}^{n}_{h},\mathbf{u}^{n}_{h})+a_{\mu}(\mu^{n},\mu^{n})+s_{h}(p_{h}^{n},p_{h}^{n})\right)\leq K, (4.10)

for all N=1,2,N=1,2,\dots, with K=Γ(ρ0|𝐮h0|2+σγϵf0(ch0))𝑑s+ac(ch0,ch0)K=\int_{\Gamma}\left(\rho^{0}|\mathbf{u}^{0}_{h}|^{2}+\frac{\sigma_{\gamma}}{\epsilon}f_{0}(c^{0}_{h})\right)ds+a_{c}(c^{0}_{h},c^{0}_{h}).

Proof:

We use induction in NN to prove (4.10). For N=0N=0, the estimate is trivial and provided by the initial condition. For the induction step, assume that it holds with N=nN=n.

Letting vh=μhn+1v_{h}=\mu^{n+1}_{h} in (3.13) and gh=[ch]tn+1g_{h}=-\left[c_{h}\right]_{t}^{n+1} in (3.14) and adding the two equations together, we get

(chn+1𝐮hn,Γμhn+1)+aμ(μhn+1,μhn+1)+(1ϵ[ch]tn+1,f0(chn))+12Δt(ac(chn+1,chn+1)ac(chn,chn)+|Δt|2ac([ch]tn+1,[ch]tn+1))+γcΔtϵ[ch]tn+12=0.-(c^{n+1}_{h}\mathbf{u}^{n}_{h},\nabla_{\Gamma}\mu^{n+1}_{h})+a_{\mu}(\mu^{n+1}_{h},\mu^{n+1}_{h})+\left(\frac{1}{\epsilon}\left[c_{h}\right]_{t}^{n+1},{f}_{0}^{\prime}(c^{n}_{h})\right)\\ +\frac{1}{2\Delta t}\left(a_{c}(c^{n+1}_{h},c^{n+1}_{h})-a_{c}(c^{n}_{h},c^{n}_{h})+|\Delta t|^{2}a_{c}(\left[c_{h}\right]_{t}^{n+1},\left[c_{h}\right]_{t}^{n+1})\right)+\frac{\gamma_{c}\Delta t}{\epsilon}\|\left[c_{h}\right]_{t}^{n+1}\|^{2}=0. (4.11)

With the truncated Taylor expansion f0(chn+1)=f0(chn)+|Δt|[ch]tn+1f0(chn)+|Δt|2|[ch]tn+1|2f0′′(ξn),f_{0}(c^{n+1}_{h})=f_{0}(c^{n}_{h})+|\Delta t|\left[c_{h}\right]_{t}^{n+1}{f}_{0}^{\prime}(c^{n}_{h})+|\Delta t|^{2}|\left[c_{h}\right]_{t}^{n+1}|^{2}{f}_{0}^{\prime\prime}(\xi^{n}), and (4.8) we get:

(1ϵ[ch]tn+1,f0(chn))=1ϵΓf0(chn+1)f0(chn)Δt𝑑sΔt2ϵΓ|[ch]tn+1|2f0′′(ξn)𝑑s1ϵΓf0(chn+1)f0(chn)Δt𝑑sLΔt2ϵ[ch]tn+12.\begin{split}\left(\frac{1}{\epsilon}\left[c_{h}\right]_{t}^{n+1},{f}_{0}^{\prime}(c^{n}_{h})\right)&=\frac{1}{\epsilon}\int_{\Gamma}\frac{f_{0}(c^{n+1}_{h})-f_{0}(c^{n}_{h})}{\Delta t}\,ds-\frac{\Delta t}{2\epsilon}\int_{\Gamma}|\left[c_{h}\right]_{t}^{n+1}|^{2}{f}_{0}^{\prime\prime}(\xi^{n})ds\\ &\geq\frac{1}{\epsilon}\int_{\Gamma}\frac{f_{0}(c^{n+1}_{h})-f_{0}(c^{n}_{h})}{\Delta t}\,ds-\frac{L\Delta t}{2\epsilon}\|\left[c_{h}\right]_{t}^{n+1}\|^{2}.\end{split}

Since the second term at the right-hand side has a negative sign, we let γc\gamma_{c} be large enough in order to cancel it with the 12\frac{1}{2} of the last term on the left-hand side of (4.11). We obtain

aμ(μhn+1,μhn+1)+1ϵΓf0(chn+1)f0(chn)Δt𝑑s+12Δt(ac(chn+1,chn+1)ac(chn,chn))+Δt(12ac([ch]tn+1,[ch]tn+1)+γc2ϵ[ch]tn+12)(chn+1𝐮hn,Γμhn+1).a_{\mu}(\mu^{n+1}_{h},\mu^{n+1}_{h})+\frac{1}{\epsilon}\int_{\Gamma}\frac{f_{0}(c^{n+1}_{h})-f_{0}(c^{n}_{h})}{\Delta t}\,ds+\frac{1}{2\Delta t}\left(a_{c}(c^{n+1}_{h},c^{n+1}_{h})-a_{c}(c^{n}_{h},c^{n}_{h})\right)\\ +\Delta t\left(\frac{1}{2}a_{c}(\left[c_{h}\right]_{t}^{n+1},\left[c_{h}\right]_{t}^{n+1})+\frac{\gamma_{c}}{2\epsilon}\|\left[c_{h}\right]_{t}^{n+1}\|^{2}\right)\leq(c^{n+1}_{h}\mathbf{u}^{n}_{h},\nabla_{\Gamma}\mu^{n+1}_{h}). (4.12)

After re-arranging terms, multiplying by Δt\Delta t, and dropping some non-negative terms on the left-hand side, we obtain the following inequality

1ϵΓf0(chn+1)𝑑s+Δtaμ(μhn+1,μhn+1)+12ac(chn+1,chn+1)+Δt2γc2ϵ[ch]tn+121ϵΓf0(chn)𝑑s+12ac(chn,chn)+Δt(chn+1𝐮hn,Γμhn+1).\frac{1}{\epsilon}\int_{\Gamma}f_{0}(c^{n+1}_{h})ds+\Delta ta_{\mu}(\mu^{n+1}_{h},\mu^{n+1}_{h})+\frac{1}{2}a_{c}(c^{n+1}_{h},c^{n+1}_{h})+\Delta t^{2}\frac{\gamma_{c}}{2\epsilon}\|\left[c_{h}\right]_{t}^{n+1}\|^{2}\\ \leq\frac{1}{\epsilon}\int_{\Gamma}f_{0}(c^{n}_{h})ds+\frac{1}{2}a_{c}(c^{n}_{h},c^{n}_{h})+\Delta t(c^{n+1}_{h}\mathbf{u}^{n}_{h},\nabla_{\Gamma}\mu^{n+1}_{h}). (4.13)

The first two terms on the right-hand side of (4.13) are bounded due to the induction assumption. To handle the third term on the right-hand side, we let c0=|Γ|1Γchn+1𝑑sc_{0}=|\Gamma|^{-1}\int_{\Gamma}c_{h}^{n+1}\,ds and use Lemma 4.1 that yields chn+1c0L(Γ)|lnh|12ϵ12|ac(chn+1,chn+1)|12\|c^{n+1}_{h}-c_{0}\|_{L^{\infty}(\Gamma)}\lesssim|\ln h|^{\frac{1}{2}}\epsilon^{-\frac{1}{2}}|a_{c}(c^{n+1}_{h},c^{n+1}_{h})|^{\frac{1}{2}} by the definition of the aca_{c} form in (3.4). This, the Cauchy inequality, and the equality 𝐮hnΓμhn+1=𝐮¯hnΓμhn+1\mathbf{u}^{n}_{h}\cdot\nabla_{\Gamma}\mu^{n+1}_{h}=\overline{\mathbf{u}}_{h}^{n}\cdot\nabla_{\Gamma}\mu^{n+1}_{h} help us with the following estimate:

Δt|(chn+1𝐮hn,Γμhn+1)|Δt|((chn+1c0)𝐮hn,Γμhn+1)|+Δt|(c0𝐮hn,Γμhn+1)|Δt𝐮¯hnchn+1c0L(Γ)Γμhn+1+Δt𝐮¯hn|c0|Γμhn+1CΔtϵ𝐮¯hn|lnh|12|ac(chn+1,chn+1)|12Γμhn+1+Δt𝐮¯hn|c0|Γμhn+1Δt|aμ(μhn+1,μhn+1)|12Kρ2M(C|lnh|12ϵ|ac(chn+1,chn+1)|12+|c0|).\begin{split}\Delta t|(c^{n+1}_{h}&\mathbf{u}^{n}_{h},\nabla_{\Gamma}\mu^{n+1}_{h})|\leq\Delta t|(\,(c^{n+1}_{h}-c_{0})\mathbf{u}^{n}_{h},\nabla_{\Gamma}\mu^{n+1}_{h})|+\Delta t|(c_{0}\mathbf{u}^{n}_{h},\nabla_{\Gamma}\mu^{n+1}_{h})|\\ &\leq\Delta t\|\overline{\mathbf{u}}_{h}^{n}\|\|c^{n+1}_{h}-c_{0}\|_{L^{\infty}(\Gamma)}\|\nabla_{\Gamma}\mu^{n+1}_{h}\|+\Delta t\|\overline{\mathbf{u}}_{h}^{n}\||c_{0}|\|\nabla_{\Gamma}\mu^{n+1}_{h}\|\\ &\leq\frac{C\Delta t}{\sqrt{\epsilon}}\|\overline{\mathbf{u}}_{h}^{n}\||\ln h|^{\frac{1}{2}}|a_{c}(c^{n+1}_{h},c^{n+1}_{h})|^{\frac{1}{2}}\|\nabla_{\Gamma}\mu^{n+1}_{h}\|+\Delta t\|\overline{\mathbf{u}}_{h}^{n}\||c_{0}|\|\nabla_{\Gamma}\mu^{n+1}_{h}\|\\ &\leq\Delta t|a_{\mu}(\mu^{n+1}_{h},\mu^{n+1}_{h})|^{\frac{1}{2}}\sqrt{\frac{K}{\rho_{2}M}}\left(\frac{C|\ln h|^{\frac{1}{2}}}{\sqrt{\epsilon}}|a_{c}(c^{n+1}_{h},c^{n+1}_{h})|^{\frac{1}{2}}+|c_{0}|\right).\end{split} (4.14)

In last inequality in (4.14), we used the induction assumption to estimate 𝐮hn\|\mathbf{u}^{n}_{h}\| and the fact that ρnρ2\rho^{n}\geq\rho_{2} by definition of the ρ(c)\rho(c) function. Letting vh=1v_{h}=1 in (3.13) we have

Γ[ch]tn+1𝑑s=0c0=|Γ|1Γch0𝑑s.\int_{\Gamma}\left[c_{h}\right]_{t}^{n+1}\,ds=0\quad\Rightarrow\quad c_{0}=|\Gamma|^{-1}\int_{\Gamma}c_{h}^{0}\,ds.

We conclude that |c0||c_{0}| in (4.14) can be bounded by a constant depending only on the initial data. Then, using Young’s inequality in (4.14) we get the following bound

Δt|(𝐮hnΓμhn+1,chn+1)|14ac(chn+1,chn+1)+|Δt|2C1|lnh|ϵaμ(μhn+1,μhn+1)+C2,\Delta t|(\mathbf{u}^{n}_{h}\cdot\nabla_{\Gamma}\mu^{n+1}_{h},c^{n+1}_{h})|\leq\frac{1}{4}a_{c}(c^{n+1}_{h},c^{n+1}_{h})+|\Delta t|^{2}\frac{C_{1}|\ln h|}{\epsilon}a_{\mu}(\mu^{n+1}_{h},\mu^{n+1}_{h})+C_{2},

with some constants C1C_{1}, C2C_{2} independent of hh, Δt\Delta t, ϵ\epsilon, and position of Γ\Gamma in the background mesh. Using this back in (4.13) with Δt\Delta t satisfying assumptions of the theorem, and applying (4.10) for N=nN=n for the remainder terms on the right-hand side of (4.13), we get:

1ϵΓf0(chn+1)𝑑s+Δt2aμ(μhn+1,μhn+1)+14ac(chn+1,chn+1)+Δt2γc2ϵ[ch]tn+12C,\frac{1}{\epsilon}\int_{\Gamma}f_{0}(c^{n+1}_{h})ds+\frac{\Delta t}{2}a_{\mu}(\mu^{n+1}_{h},\mu^{n+1}_{h})+\frac{1}{4}a_{c}(c^{n+1}_{h},c^{n+1}_{h})+\frac{\Delta t^{2}\gamma_{c}}{2\epsilon}\|\left[c_{h}\right]_{t}^{n+1}\|^{2}\leq C, (4.15)

with some C>0C>0 independent of hh, Δt\Delta t, ϵ\epsilon and the position of Γ\Gamma in the mesh. We need this bound later in the proof.

Let 𝐯h=𝐮hn+1\mathbf{v}_{h}=\mathbf{u}^{n+1}_{h} in (3.15) and qh=phq_{h}=-p_{h} in (3.16), and add the two equations together. We also apply Lemma 3.1 to handle the cc-form. This brings us to the following equality:

12Δt((ρn)12𝐮¯hn+12(ρn)12𝐮¯hn2+|Δt|2(ρn)12[𝐮¯h]tn+12)\displaystyle\frac{1}{2\Delta t}\left(\|(\rho^{n})^{\frac{1}{2}}\overline{\mathbf{u}}_{h}^{n+1}\|^{2}-\|(\rho^{n})^{\frac{1}{2}}\overline{\mathbf{u}}_{h}^{n}\|^{2}+|\Delta t|^{2}\|(\rho^{n})^{\frac{1}{2}}\left[\overline{\mathbf{u}}_{h}\right]_{t}^{n+1}\|^{2}\right)
12Γ|θn+1|2divΓ(chn+1𝐮¯hn)|𝐮¯hn+1|2ds+a(ηn+1;𝐮hn+1,𝐮hn+1)+s(phn+1,phn+1)\displaystyle\quad-\frac{1}{2}\int_{\Gamma}|\theta^{n+1}|^{2}{\mathop{\,\rm div}}_{\Gamma}(c_{h}^{n+1}\overline{\mathbf{u}}_{h}^{n})|\overline{\mathbf{u}}_{h}^{n+1}|^{2}\,ds+a(\eta^{n+1};\mathbf{u}^{n+1}_{h},\mathbf{u}^{n+1}_{h})+s(p_{h}^{n+1},p_{h}^{n+1})
=(σγchn+1Γμhn+1,𝐮hn+1)+M((Γ(θn+1𝐮¯hn+1))Γμhn+1,θn+1𝐮hn+1).\displaystyle\quad=-(\sigma_{\gamma}c^{n+1}_{h}\nabla_{\Gamma}\mu^{n+1}_{h},\mathbf{u}^{n+1}_{h})+M\left((\nabla_{\Gamma}(\theta^{n+1}\overline{\mathbf{u}}_{h}^{n+1}))\nabla_{\Gamma}\mu^{n+1}_{h},\theta^{n+1}\mathbf{u}^{n+1}_{h}\right).

By adding ±12Δt(ρn+1)12𝐮¯hn+12\pm\frac{1}{2\Delta t}\|(\rho^{n+1})^{\frac{1}{2}}\overline{\mathbf{u}}_{h}^{n+1}\|^{2} and re-arranging terms, we get

12Δt((ρn+1)12𝐮¯hn+12(ρn)12𝐮¯hn2+|Δt|2(ρn)12[𝐮¯h]tn+12)\displaystyle\frac{1}{2\Delta t}\left(\|(\rho^{n+1})^{\frac{1}{2}}\overline{\mathbf{u}}_{h}^{n+1}\|^{2}-\|(\rho^{n})^{\frac{1}{2}}\overline{\mathbf{u}}_{h}^{n}\|^{2}+|\Delta t|^{2}\|(\rho^{n})^{\frac{1}{2}}\left[\overline{\mathbf{u}}_{h}\right]_{t}^{n+1}\|^{2}\right)
12Γρn+1ρnΔt|𝐮¯hn+1|2𝑑s12Γ|θn+1|2divΓ(chn+1𝐮¯hn)|𝐮¯hn+1|2ds\displaystyle\quad-\frac{1}{2}\int_{\Gamma}\frac{\rho^{n+1}-\rho^{n}}{\Delta t}|\overline{\mathbf{u}}_{h}^{n+1}|^{2}\,ds-\frac{1}{2}\int_{\Gamma}|\theta^{n+1}|^{2}{\mathop{\,\rm div}}_{\Gamma}(c_{h}^{n+1}\overline{\mathbf{u}}_{h}^{n})|\overline{\mathbf{u}}_{h}^{n+1}|^{2}\,ds
+a(ηn+1;𝐮hn+1,𝐮hn+1)+s(phn+1,phn+1)\displaystyle\quad+a(\eta^{n+1};\mathbf{u}^{n+1}_{h},\mathbf{u}^{n+1}_{h})+s(p_{h}^{n+1},p_{h}^{n+1})
=(σγchn+1Γμhn+1,𝐮hn+1)+M((Γ(θn+1𝐮¯hn+1))Γμhn+1,θn+1𝐮hn+1).\displaystyle\quad=-(\sigma_{\gamma}c^{n+1}_{h}\nabla_{\Gamma}\mu^{n+1}_{h},\mathbf{u}^{n+1}_{h})+M\left((\nabla_{\Gamma}(\theta^{n+1}\overline{\mathbf{u}}_{h}^{n+1}))\nabla_{\Gamma}\mu^{n+1}_{h},\theta^{n+1}\mathbf{u}^{n+1}_{h}\right). (4.16)

Denote by 𝒫h:H1(Γ)Vhk\mathcal{P}_{h}\,:\,H^{1}(\Gamma)\to V^{k}_{h} an H1H^{1}-type projection into VhkV^{k}_{h} given by the aμa_{\mu} bilinear form:

aμ(𝒫h(u),vh)=M(Γu,Γvh)vhVhk.a_{\mu}(\mathcal{P}_{h}(u),v_{h})=M\left(\nabla_{\Gamma}u,\nabla_{\Gamma}v_{h}\right)\quad\forall\,v_{h}\in V_{h}^{k}.

By standard analysis of the TraceFEM for the Laplace-Beltrami problem, e.g. [37], we have

u𝒫h(u)hΓu.\|u-\mathcal{P}_{h}(u)\|\lesssim h\|\nabla_{\Gamma}u\|. (4.17)

Since ρ\rho is a convex function of cc and dρdc0\frac{d\rho}{dc}\geq 0, we have

ρn+1ρndρdc|chn+1(chn+1chn)=|θn+1|2(chn+1chn).\rho^{n+1}-\rho^{n}\leq\frac{d\rho}{dc}\Big{|}_{c^{n+1}_{h}}(c^{n+1}_{h}-c^{n}_{h})=|\theta^{n+1}|^{2}(c^{n+1}_{h}-c^{n}_{h}). (4.18)

Using (4.18) and eq. (3.13) with vh=𝒫h(|θn+1𝐮¯hn+1|2)v_{h}=\mathcal{P}_{h}\left(|\theta^{n+1}\overline{\mathbf{u}}_{h}^{n+1}|^{2}\right), for the terms on the second line of (4.16) we obtain

Γρn+1ρnΔt|𝐮¯hn+1|2+|θn+1|2divΓ(chn+1𝐮¯hn)|𝐮¯hn+1|2ds\displaystyle\int_{\Gamma}\frac{\rho^{n+1}-\rho^{n}}{\Delta t}|\overline{\mathbf{u}}_{h}^{n+1}|^{2}+|\theta^{n+1}|^{2}{\mathop{\,\rm div}}_{\Gamma}(c_{h}^{n+1}\overline{\mathbf{u}}_{h}^{n})|\overline{\mathbf{u}}_{h}^{n+1}|^{2}\,ds
Γ(cn+1cnΔt+divΓ(chn+1𝐮¯hn))|θn+1𝐮¯hn+1|2𝑑s\displaystyle\quad\leq\int_{\Gamma}\left(\frac{c^{n+1}-c^{n}}{\Delta t}+{\mathop{\,\rm div}}_{\Gamma}(c_{h}^{n+1}\overline{\mathbf{u}}_{h}^{n})\right)|\theta^{n+1}\overline{\mathbf{u}}_{h}^{n+1}|^{2}\,ds
=ΓMΓμhn+1Γ|θn+1𝐮¯hn+1|2ds+I(eh),\displaystyle\quad=-\int_{\Gamma}M\nabla_{\Gamma}\mu^{n+1}_{h}\cdot\nabla_{\Gamma}|\theta^{n+1}\overline{\mathbf{u}}_{h}^{n+1}|^{2}\,ds+I(e_{h}), (4.19)

with

I(eh)=Γ(chn+1chnΔt+divΓ(chn+1𝐮¯hn))eh𝑑s,eh|θn+1𝐮¯hn+1|2𝒫h(|θn+1𝐮¯hn+1|2).I(e_{h})=\int_{\Gamma}\left(\frac{c^{n+1}_{h}-c^{n}_{h}}{\Delta t}+{\mathop{\,\rm div}}_{\Gamma}(c_{h}^{n+1}\overline{\mathbf{u}}_{h}^{n})\right)e_{h}\,ds,\quad e_{h}\coloneqq|\theta^{n+1}\overline{\mathbf{u}}_{h}^{n+1}|^{2}-\mathcal{P}_{h}\left(|\theta^{n+1}\overline{\mathbf{u}}_{h}^{n+1}|^{2}\right).

Next, we use (4.19) to simplify (4.16) as follows:

12Δt((ρn+1)12𝐮¯hn+12(ρn)12𝐮¯hn2+|Δt|2(ρn)12[𝐮¯h]tn+12)+a(ηn+1;𝐮hn+1,𝐮hn+1)+s(phn+1,phn+1)(σγchn+1Γμhn+1,𝐮hn+1)+I(eh).\frac{1}{2\Delta t}\left(\|(\rho^{n+1})^{\frac{1}{2}}\overline{\mathbf{u}}_{h}^{n+1}\|^{2}-\|(\rho^{n})^{\frac{1}{2}}\overline{\mathbf{u}}_{h}^{n}\|^{2}+|\Delta t|^{2}\|(\rho^{n})^{\frac{1}{2}}\left[\overline{\mathbf{u}}_{h}\right]_{t}^{n+1}\|^{2}\right)\\ +a(\eta^{n+1};\mathbf{u}^{n+1}_{h},\mathbf{u}^{n+1}_{h})+s(p_{h}^{n+1},p_{h}^{n+1})\leq-(\sigma_{\gamma}c^{n+1}_{h}\nabla_{\Gamma}\mu^{n+1}_{h},\mathbf{u}^{n+1}_{h})+I(e_{h}). (4.20)

After adding (4.12) multiplied by σγ\sigma_{\gamma} to (4.20) and dropping some non-negative terms on the left-hand side, we arrive at

12Δt((ρn+1)12𝐮¯hn+12(ρn)12𝐮¯hn2+|Δt|2(ρn)12[𝐮¯h]tn+12\displaystyle\frac{1}{2\Delta t}\left(\|(\rho^{n+1})^{\frac{1}{2}}\overline{\mathbf{u}}_{h}^{n+1}\|^{2}-\|(\rho^{n})^{\frac{1}{2}}\overline{\mathbf{u}}_{h}^{n}\|^{2}+|\Delta t|^{2}\|(\rho^{n})^{\frac{1}{2}}\left[\overline{\mathbf{u}}_{h}\right]_{t}^{n+1}\|^{2}\right.
+σγ(a(chn+1,chn+1)a(chn,chn))+2σγϵΓ(f0(chn+1)f0(chn)))ds+sh(phn+1,phn+1)\displaystyle\quad\left.+\sigma_{\gamma}(a(c^{n+1}_{h},c^{n+1}_{h})-a(c^{n}_{h},c^{n}_{h}))+\frac{2\sigma_{\gamma}}{\epsilon}\int_{\Gamma}(f_{0}(c^{n+1}_{h})-f_{0}(c^{n}_{h}))\right)ds+s_{h}(p_{h}^{n+1},p_{h}^{n+1})
+a(ηn+1;𝐮hn+1,𝐮hn+1)+σγaμ(μhn+1,μhn+1)\displaystyle\quad+a(\eta^{n+1};\mathbf{u}^{n+1}_{h},\mathbf{u}^{n+1}_{h})+\sigma_{\gamma}a_{\mu}(\mu^{n+1}_{h},\mu^{n+1}_{h})
Δt(σγchn+1Γμhn+1,[𝐮h]tn+1)+I(eh).\displaystyle\quad\leq-\Delta t(\sigma_{\gamma}c^{n+1}_{h}\nabla_{\Gamma}\mu^{n+1}_{h},\left[\mathbf{u}_{h}\right]_{t}^{n+1})+I(e_{h}). (4.21)

It remains to estimate the terms on the right-hand side of (4.21). We handle the first term by invoking the result of Lemma 4.1 as follows:

Δt|(σγ\displaystyle\Delta t|(\sigma_{\gamma} chn+1Γμhn+1,[𝐮h]tn+1)|σγΔt|((chn+1c0)Γμhn+1,[𝐮h]tn+1)|+σγΔt|(c0Γμhn+1,[𝐮h]tn+1)|\displaystyle c^{n+1}_{h}\nabla_{\Gamma}\mu^{n+1}_{h},\left[\mathbf{u}_{h}\right]_{t}^{n+1})|\leq\sigma_{\gamma}\Delta t|((c^{n+1}_{h}-c_{0})\nabla_{\Gamma}\mu^{n+1}_{h},\left[\mathbf{u}_{h}\right]_{t}^{n+1})|+\sigma_{\gamma}\Delta t|(c_{0}\nabla_{\Gamma}\mu^{n+1}_{h},\left[\mathbf{u}_{h}\right]_{t}^{n+1})|
σγΔtchn+1c0L(Γ)Γμhn+1[𝐮¯h]tn+1+σγΔt|c0|Γμhn+1[𝐮¯h]tn+1\displaystyle\leq\sigma_{\gamma}\Delta t\|c^{n+1}_{h}-c_{0}\|_{L^{\infty}(\Gamma)}\|\nabla_{\Gamma}\mu^{n+1}_{h}\|\|\left[\overline{\mathbf{u}}_{h}\right]_{t}^{n+1}\|+\sigma_{\gamma}\Delta t|c_{0}|\|\nabla_{\Gamma}\mu^{n+1}_{h}\|\|\left[\overline{\mathbf{u}}_{h}\right]_{t}^{n+1}\|
Cρ212ϵ12|lnh|12σγΔt|ac(chn+1,chn+1)|12Γμhn+1(ρn)12[𝐮¯h]tn+1\displaystyle\leq\frac{C}{\rho_{2}^{\frac{1}{2}}\epsilon^{\frac{1}{2}}}|\ln h|^{\frac{1}{2}}\sigma_{\gamma}\Delta t|a_{c}(c^{n+1}_{h},c^{n+1}_{h})|^{\frac{1}{2}}\|\nabla_{\Gamma}\mu^{n+1}_{h}\|\|(\rho^{n})^{\frac{1}{2}}\left[\overline{\mathbf{u}}_{h}\right]_{t}^{n+1}\|
+σγΔtρ212|c0|Γμhn+1(ρn)12[𝐮¯h]tn+1.\displaystyle\qquad+\frac{\sigma_{\gamma}\Delta t}{\rho_{2}^{\frac{1}{2}}}|c_{0}|\|\nabla_{\Gamma}\mu^{n+1}_{h}\|\|(\rho^{n})^{\frac{1}{2}}\left[\overline{\mathbf{u}}_{h}\right]_{t}^{n+1}\|. (4.22)

Thanks to the a priori bound ac(chn+1,chn+1)Ca_{c}(c^{n+1}_{h},c^{n+1}_{h})\leq C from (4.15), estimate (4.22) yields

Δt|(σγchn+1Γμhn+1,[𝐮h]tn+1)|CσγΔt(ϵ12|lnh|12+1)(ρn)12[𝐮¯h]tn+1|aμ(μhn+1,μhn+1)|12.\Delta t|(\sigma_{\gamma}c^{n+1}_{h}\nabla_{\Gamma}\mu^{n+1}_{h},\left[\mathbf{u}_{h}\right]_{t}^{n+1})|\leq C\sigma_{\gamma}\Delta t({\epsilon^{-\frac{1}{2}}}|\ln h|^{\frac{1}{2}}+1)\|(\rho^{n})^{\frac{1}{2}}\left[\overline{\mathbf{u}}_{h}\right]_{t}^{n+1}\||a_{\mu}(\mu^{n+1}_{h},\mu^{n+1}_{h})|^{\frac{1}{2}}. (4.23)

With the help of ΔtC|lnh|1ϵ\Delta t\leq C|\ln h|^{-1}\epsilon for sufficiently small CC and ϵ1\epsilon\lesssim 1 we obtain

Δt|(σγchn+1Γμhn+1,[𝐮h]tn+1)|\displaystyle\Delta t|(\sigma_{\gamma}c^{n+1}_{h}\nabla_{\Gamma}\mu^{n+1}_{h},\left[\mathbf{u}_{h}\right]_{t}^{n+1})| Δt2(ρn)12[𝐮¯h]tn+12+CσγΔt(ϵ12|lnh|12+1)2aμ(μhn+1,μhn+1)\displaystyle\leq\frac{\Delta t}{2}\|(\rho^{n})^{\frac{1}{2}}\left[\overline{\mathbf{u}}_{h}\right]_{t}^{n+1}\|^{2}+C\sigma_{\gamma}\Delta t({\epsilon^{-\frac{1}{2}}}|\ln h|^{\frac{1}{2}}+1)^{2}a_{\mu}(\mu^{n+1}_{h},\mu^{n+1}_{h})
Δt2(ρn)12[𝐮¯h]tn+12+12σγaμ(μhn+1,μhn+1).\displaystyle\leq\frac{\Delta t}{2}\|(\rho^{n})^{\frac{1}{2}}\left[\overline{\mathbf{u}}_{h}\right]_{t}^{n+1}\|^{2}+\frac{1}{2}\sigma_{\gamma}a_{\mu}(\mu^{n+1}_{h},\mu^{n+1}_{h}). (4.24)

Now, we proceed with the terms in IhI_{h}. For the first terms in IhI_{h}, the Cauchy-Schwarz inequality and estimate (4.17) for the L2(Γ)L^{2}(\Gamma) norm of ehe_{h} gives:

Γ(chn+1chnΔt)eh𝑑s[ch]tn+1ehh[ch]tn+1Γ(|θn+1𝐮¯hn+1|2)\int_{\Gamma}\left(\frac{c_{h}^{n+1}-c_{h}^{n}}{\Delta t}\right)e_{h}\,ds\leq\|\left[c_{h}\right]_{t}^{n+1}\|\|e_{h}\|\lesssim h\|\left[c_{h}\right]_{t}^{n+1}\|\|\nabla_{\Gamma}(|\theta^{n+1}\overline{\mathbf{u}}_{h}^{n+1}|^{2})\| (4.25)

To estimate the right-most factor, we need the inequalities

𝐮¯hn+1L(Γ)2|lnh|𝐮¯hn+1H1(Γ)2+h1𝐧|𝐮¯hn+1L2(ΩhΓ)2|lnh|a(ηn+1,𝐮hn+1,𝐮hn+1),\|\overline{\mathbf{u}}_{h}^{n+1}\|^{2}_{L^{\infty}(\Gamma)}\lesssim|\ln h|\|\overline{\mathbf{u}}_{h}^{n+1}\|^{2}_{H^{1}(\Gamma)}+h^{-1}\|\mathbf{n}\cdot\nabla|\overline{\mathbf{u}}_{h}^{n+1}\|^{2}_{L^{2}(\Omega_{h}^{\Gamma})}\lesssim|\ln h|a(\eta^{n+1},\mathbf{u}_{h}^{n+1},\mathbf{u}_{h}^{n+1}),

which follow by applying Lemma 4.1 componentwise and then using the Korn inequality (2.12) and the fact that ηn+1\eta^{n+1} is uniformly bounded from below (see the definition in (3.5)). Thanks to 0θ210\leq\theta^{2}\lesssim 1 and 0dθ2dc=d2ρdc210\leq\frac{d\theta^{2}}{dc}=\frac{d^{2}\rho}{dc^{2}}\lesssim 1, we have

Γ(|θn+1𝐮¯hn+1|2)\displaystyle\|\nabla_{\Gamma}(|\theta^{n+1}\overline{\mathbf{u}}_{h}^{n+1}|^{2})\| |𝐮¯hn+1|2Γchn+1+Γ|𝐮¯hn+1|2\displaystyle\lesssim\||\overline{\mathbf{u}}_{h}^{n+1}|^{2}\nabla_{\Gamma}c^{n+1}_{h}\|+\|\nabla_{\Gamma}|\overline{\mathbf{u}}_{h}^{n+1}|^{2}\|
𝐮¯hn+1L(Γ)2Γchn+1+𝐮¯hn+1L(Γ)Γ𝐮¯hn+1\displaystyle\lesssim\|\overline{\mathbf{u}}_{h}^{n+1}\|^{2}_{L^{\infty}(\Gamma)}\|\nabla_{\Gamma}c^{n+1}_{h}\|+\|\overline{\mathbf{u}}_{h}^{n+1}\|_{L^{\infty}(\Gamma)}\|\nabla_{\Gamma}\overline{\mathbf{u}}_{h}^{n+1}\|
|lnh|a(ηn+1,𝐮hn+1,𝐮hn+1)Γchn+1+|lnh|12a(ηn+1,𝐮hn+1,𝐮hn+1)\displaystyle\lesssim|\ln h|a(\eta^{n+1},\mathbf{u}_{h}^{n+1},\mathbf{u}_{h}^{n+1})\|\nabla_{\Gamma}c^{n+1}_{h}\|+|\ln h|^{\frac{1}{2}}a(\eta^{n+1},\mathbf{u}_{h}^{n+1},\mathbf{u}_{h}^{n+1})
(|lnh|ϵ12+|lnh|12)a(ηn+1,𝐮hn+1,𝐮hn+1)\displaystyle\lesssim(|\ln h|\epsilon^{-\frac{1}{2}}+|\ln h|^{\frac{1}{2}})a(\eta^{n+1},\mathbf{u}_{h}^{n+1},\mathbf{u}_{h}^{n+1})
|lnh|ϵ12a(ηn+1,𝐮hn+1,𝐮hn+1).\displaystyle\lesssim|\ln h|\epsilon^{-\frac{1}{2}}a(\eta^{n+1},\mathbf{u}_{h}^{n+1},\mathbf{u}_{h}^{n+1}).

Using this together with the a priori bound for [ch]tn+1\|\left[c_{h}\right]_{t}^{n+1}\| from (4.15) in (4.25) and using the assumption hc|lnh|1Δth\leq c|\ln h|^{-1}\Delta t for sufficiently small cc, we have

Γ(chn+1chnΔt)eh𝑑sch|lnh|Δt1a(ηn+1,𝐮hn+1,𝐮hn+1)14a(ηn+1,𝐮hn+1,𝐮hn+1).\int_{\Gamma}\left(\frac{c_{h}^{n+1}-c_{h}^{n}}{\Delta t}\right)e_{h}\,ds\leq c\,h|\ln h|\Delta t^{-1}a(\eta^{n+1},\mathbf{u}_{h}^{n+1},\mathbf{u}_{h}^{n+1})\leq\frac{1}{4}a(\eta^{n+1},\mathbf{u}_{h}^{n+1},\mathbf{u}_{h}^{n+1}). (4.26)

Next, we treat the second term in I(eh)I(e_{h}). Using Cauchy–Schwarz inequality and estimating eh\|e_{h}\| as above we have

ΓdivΓ(chn+1𝐮¯hn)ehdsh|lnh|ϵ12divΓ(chn+1𝐮¯hn)a(ηn+1,𝐮hn+1,𝐮hn+1).\int_{\Gamma}{\mathop{\,\rm div}}_{\Gamma}(c_{h}^{n+1}\overline{\mathbf{u}}_{h}^{n})e_{h}\,ds\lesssim h|\ln h|\epsilon^{-\frac{1}{2}}\|{\mathop{\,\rm div}}_{\Gamma}(c_{h}^{n+1}\overline{\mathbf{u}}_{h}^{n})\|a(\eta^{n+1},\mathbf{u}_{h}^{n+1},\mathbf{u}_{h}^{n+1}). (4.27)

By the triangle inequality

divΓ(chn+1𝐮¯hn)chn+1divΓ(𝐮¯hn)+𝐮hnΓchn+1.\|{\mathop{\,\rm div}}_{\Gamma}(c_{h}^{n+1}\overline{\mathbf{u}}_{h}^{n})\|\leq\|c_{h}^{n+1}{\mathop{\,\rm div}}_{\Gamma}(\overline{\mathbf{u}}_{h}^{n})\|+\|\mathbf{u}^{n}_{h}\cdot\nabla_{\Gamma}c_{h}^{n+1}\|.

Each term on the right hand side can be treated individually by invoking Lemma 4.1, a priori bound from (4.15), and induction assumption:

chn+1divΓ(𝐮¯hn)chn+1L(Γ)divΓ(𝐮¯hn)|lnh|12chn+1H1(Γ)divΓ(𝐮¯hn)|lnh|12ϵ12divΓ(𝐮¯hn)|lnh|12ϵ12|a(ηn,𝐮hn,𝐮hn)|12|lnh|12ϵ12|Δt|12.\begin{split}\|c_{h}^{n+1}{\mathop{\,\rm div}}_{\Gamma}(\overline{\mathbf{u}}_{h}^{n})\|&\leq\|c_{h}^{n+1}\|_{L^{\infty}(\Gamma)}\|{\mathop{\,\rm div}}_{\Gamma}(\overline{\mathbf{u}}_{h}^{n})\|\\ &\lesssim|\ln h|^{\frac{1}{2}}\|c_{h}^{n+1}\|_{H^{1}(\Gamma)}\|{\mathop{\,\rm div}}_{\Gamma}(\overline{\mathbf{u}}_{h}^{n})\|\\ &\lesssim|\ln h|^{\frac{1}{2}}\epsilon^{-\frac{1}{2}}\|{\mathop{\,\rm div}}_{\Gamma}(\overline{\mathbf{u}}_{h}^{n})\|\lesssim|\ln h|^{\frac{1}{2}}\epsilon^{-\frac{1}{2}}|a(\eta^{n},\mathbf{u}_{h}^{n},\mathbf{u}_{h}^{n})|^{\frac{1}{2}}\lesssim|\ln h|^{\frac{1}{2}}\epsilon^{-\frac{1}{2}}|\Delta t|^{-\frac{1}{2}}.\end{split}

and

𝐮hnΓchn+1𝐮¯hnL(Γ)Γchn+1ϵ12𝐮¯hnL(Γ)ϵ12|lnh|12𝐮¯hnH1(Γ)ϵ12|lnh|12|a(ηn,𝐮hn,𝐮hn)|12|lnh|12ϵ12|Δt|12.\begin{split}\|\mathbf{u}^{n}_{h}\cdot\nabla_{\Gamma}c_{h}^{n+1}\|&\leq\|\overline{\mathbf{u}}_{h}^{n}\|_{L^{\infty}(\Gamma)}\|\nabla_{\Gamma}c_{h}^{n+1}\|\\ &\lesssim\epsilon^{-\frac{1}{2}}\|\overline{\mathbf{u}}_{h}^{n}\|_{L^{\infty}(\Gamma)}\lesssim\epsilon^{-\frac{1}{2}}|\ln h|^{\frac{1}{2}}\|\overline{\mathbf{u}}_{h}^{n}\|_{H^{1}(\Gamma)}\\ &\lesssim\epsilon^{-\frac{1}{2}}|\ln h|^{\frac{1}{2}}|a(\eta^{n},\mathbf{u}_{h}^{n},\mathbf{u}_{h}^{n})|^{\frac{1}{2}}\lesssim|\ln h|^{\frac{1}{2}}\epsilon^{-\frac{1}{2}}|\Delta t|^{-\frac{1}{2}}.\end{split}

Using these estimates back into (4.27) and with the assumption hc|lnh|32ϵ|Δt|12h\leq c|\ln h|^{-\frac{3}{2}}\epsilon|\Delta t|^{-\frac{1}{2}} for sufficiently small cc, we arrive at the estimate

|I(eh)|12a(ηn+1;𝐮hn+1,𝐮hn+1).|I(e_{h})|\leq\frac{1}{2}a(\eta^{n+1};\mathbf{u}_{h}^{n+1},\mathbf{u}_{h}^{n+1}). (4.28)

Finally, we substitute (4.24) and (4.28) in (4.21) to obtain

12Δt((ρn+1)12𝐮¯hn+12+σγa(chn+1,chn+1)+2σγϵΓf0(chn+1)𝑑s)+12a(ηn+1;𝐮hn+1,𝐮hn+1)+12σγaμ(μhn+1,μhn+1)+sh(phn+1,phn+1)12Δt((ρn)12𝐮¯hn2+σγa(chn,chn)+2σγϵΓf0(chn)𝑑s).\frac{1}{2\Delta t}\left(\|(\rho^{n+1})^{\frac{1}{2}}\overline{\mathbf{u}}_{h}^{n+1}\|^{2}+\sigma_{\gamma}a(c^{n+1}_{h},c^{n+1}_{h})+\frac{2\sigma_{\gamma}}{\epsilon}\int_{\Gamma}f_{0}(c^{n+1}_{h})ds\right)+\frac{1}{2}a(\eta^{n+1};\mathbf{u}^{n+1}_{h},\mathbf{u}^{n+1}_{h})\\ +\frac{1}{2}\sigma_{\gamma}a_{\mu}(\mu^{n+1}_{h},\mu^{n+1}_{h})+s_{h}(p_{h}^{n+1},p_{h}^{n+1})\leq\frac{1}{2\Delta t}\left(\|(\rho^{n})^{\frac{1}{2}}\overline{\mathbf{u}}_{h}^{n}\|^{2}+\sigma_{\gamma}a(c^{n}_{h},c^{n}_{h})+\frac{2\sigma_{\gamma}}{\epsilon}\int_{\Gamma}f_{0}(c^{n}_{h})ds\right).

This completes the induction step. \square

Remark 4.3

From the proof we see that the assumption hc|lnh|1min{Δt,|lnh|12ϵ|Δt|12}h\leq c|\ln h|^{-1}\min\{\Delta t,|\ln h|^{-\frac{1}{2}}\epsilon|\Delta t|^{\frac{1}{2}}\} results from the fact that at the discrete level we cannot test the transport equation for the order parameter with vh=|𝐮hn+1|2v_{h}=|\mathbf{u}_{h}^{n+1}|^{2}, and so we have to project |𝐮hn+1|2|\mathbf{u}_{h}^{n+1}|^{2} (or |𝐮¯hn+1|2|\overline{\mathbf{u}}_{h}^{n+1}|^{2} for surfaces) in the finite element space VhkV^{k}_{h} and handle the resulting inconsistency. If the finite element velocity space is such that |𝐯h|2Vhk|\mathbf{v}_{h}|^{2}\in V_{h}^{k} for 𝐯h𝐕h\mathbf{v}_{h}\in\mathbf{V}_{h}, then the upper bound on hh is not needed in the analysis (in the surface case we still would need to handle |𝐮hn+1|2|𝐮¯hn+1|2|\mathbf{u}_{h}^{n+1}|^{2}-|\overline{\mathbf{u}}_{h}^{n+1}|^{2}, but this is possible due to the control over 𝐧𝐮hn+12\|\mathbf{n}\cdot\mathbf{u}_{h}^{n+1}\|^{2} that we have thanks to penalty term in the TraceFEM formulation). An example, when |𝐯h|2Vhk|\mathbf{v}_{h}|^{2}\in V_{h}^{k} holds, is 𝐏𝟏\bf P_{1}P1P_{1} stabilized velocity–pressure element combined with P2P_{2} element for the order parameter and chemical potential.

5 Numerical results

After checking the convergence orders of the method described in Sec. 3, we present a series of numerical tests to study well-known two-phase fluid flows (the Kelvin–Helmholtz and Rayleigh–Taylor instabilities) on surfaces. Thanks to our method, we can investigate the effect of line tension on such instabilities. In addition, for the Rayleigh–Taylor instability we assess the effect of viscosity and surface shape.

For all the tests, we choose 𝐏𝟐\bf P_{2}P1P_{1} finite elements for fluid velocity and pressure and P1P_{1}P1P_{1} finite elements for surface fraction and chemical potential.

5.1.   Convergence test

We proceed with checking the spatial accuracy of the finite element method described in Sec. 3 with a benchmark test. The aim is to validate our implementation of the method. For this purpose, we consider the two-phase fluid system on the unit sphere centered at the origin. The surface is characterized as the zero level set of function ϕ(𝐱)=𝐱21\phi(\mathbf{x})=\|\mathbf{x}\|_{2}-1 and is embedded in an outer cubic domain Ω=[5/3,5/3]3\Omega=[-5/3,5/3]^{3}. We choose Van der Waals “tanh” exact solution for the surface fraction and solenoidal Killing vector field for velocity:

c(t,𝐱)=12(1+tanhzcos(πt)ysin(πt)22ϵ),𝐮(t,𝐱)=π(0,z,y)T.c^{*}(t,\mathbf{x})=\frac{1}{2}\left(1+\tanh\frac{z\cos(\pi t)-y\sin(\pi t)}{2\sqrt{2}\epsilon}\right),\quad\mathbf{u}^{*}(t,\mathbf{x})=\pi\,(0,-z,y)^{T}.

Nonzero CH equation forcing term is computed from (2.6). We set ϵ=0.05\epsilon=0.05. Fig. 1 (leftmost panel) displays c(0,𝐱)c(0,\mathbf{x}). For this test, we consider fluids with matching densities and viscosities: ρ1=ρ2=1\rho_{1}=\rho_{2}=1 and η1=η2=1\eta_{1}=\eta_{2}=1. In addition, we set M=0.05M=0.05 and σγ=0\sigma_{\gamma}=0. The time interval of interest is t[0,1]t\in[0,1], during which the initial configuration c0c_{0} is rotated by 180180^{\circ}. See Fig. 1. Notice that by setting σγ=0\sigma_{\gamma}=0 the NSCH system one-way coupled: phase-separation is affected by the fluid flow, but not vice versa.

\begin{overpic}[width=65.04034pt,grid=false]{images/1a-min.png} \put(30.0,90.0){$t=0$} \end{overpic}
\begin{overpic}[width=65.04034pt,grid=false]{images/2a-min.png} \put(25.0,90.0){$t=0.2$} \end{overpic}
\begin{overpic}[width=65.04034pt,grid=false]{images/3a-min.png} \put(25.0,90.0){$t=0.6$} \end{overpic}
\begin{overpic}[width=65.04034pt,grid=false]{images/4a-min.png} \put(30.0,90.0){$t=1$} \end{overpic}
Figure 1: Evolution of the surface fraction cc over time computed with mesh =5\ell=5.

The initial triangulation 𝒯h\mathcal{T}_{h_{\ell}} of Ω\Omega consists of eight sub-cubes, where each of the sub-cubes is further subdivided into six tetrahedra. Further, the mesh is refined towards the surface, and \ell\in\mathbb{N} denotes the level of refinement, with the associated mesh size h=10/32+1h_{\ell}=\frac{10/3}{2^{\ell+1}}. For the purpose of numerical integration, we apply several “virtual” levels of refinements for the tetrahedra cut by the mesh and integrate our bilinear forms over a piecewise planar approximation of Γ\Gamma on this virtual grid. This allowed us to apply a standard quadrature rule and reduce the geometric error in our convergence test. The time step was refined together with the mesh size according to Δt=1/(2522)\Delta t=1/(25\cdot 2^{\ell-2}). For this test, we used BDF2 for the time discretization of the Cahn–Hilliard problem at Step 1, instead of BDF1 as reported in (3.13). This is why the time step is refined linearly.

Table 1 reports the H1H_{1} and L2L_{2} errors for the velocity and the L2L_{2} error for the order parameter at the end of the time interval, i.e. t=1t=1, for levels =3,4,5\ell=3,4,5. For each mesh, Table 1 gives the number of sublevels (virtual levels) used for more accurate numerical integration. We observe slightly better than expected convergence rates, which might be the effect of the interplay between interpolation and geometric error reduction.

\ell sublevels 𝐮𝐮hH1(Γ)||\mathbf{u}^{*}-\mathbf{u}_{h}||_{H^{1}(\Gamma)} rate 𝐮𝐮hL2(Γ)||\mathbf{u}^{*}-\mathbf{u}_{h}||_{L^{2}(\Gamma)} rate cchL2(Γ)||c^{*}-c_{h}||_{L^{2}(\Gamma)} rate
3 1 0.081485 0.010026 0.311232
4 2 0.016800 2.42 0.000619 8.20 0.081597 1.91
5 4 0.003905 2.15 0.000046 6.73 0.015086 2.70
Table 1: H1H_{1} and L2L_{2} errors for the velocity and the L2L_{2} error for surface fraction at t=1t=1 for levels =3,4,5\ell=3,4,5 and rate of convergence.

5.2.   The Kelvin–Helmholtz instability

While the Kelvin–Helmholtz (KH) instability is a classical example of two-phase fluid flow in planar or volumetric domains, the number of numerical studies on curved surfaces is limited. The KH instability arises when there is a difference in velocity at the interface between the two fluids and a perturbation is added to the interface. This perturbation eventually makes the interface curl up and generates a vortex strip. Here, we will simulate the KH instability on a sphere and investigate the effect of varying line tension.

To design this experiment, we follow what done in [26, 24, 34]. The initial velocity field is given by the counter-rotating upper and lower hemispheres with speed approximately equal 1 closer to equator and vanishing at the poles. The velocity field has a sharp transition layer along the equator, where the perturbation is added. See, e.g., [34] for details on the perturbation. The initial surface fraction is given by

c0=12(1+tanhz22ϵ),c_{0}=\frac{1}{2}\left(1+\tanh\frac{z}{2\sqrt{2}\epsilon}\right),

where ϵ=0.01\epsilon=0.01. Also for this test, we consider fluids with matching densities and viscosities: ρ1=ρ2=1\rho_{1}=\rho_{2}=1 and η1=η2=105\eta_{1}=\eta_{2}=10^{-5}. In addition, we set M=0.01M=0.01. We consider time interval [0,10][0,10].

We select mesh level =6\ell=6 (see mesh description for the previous test). We choose Δt=1/640\Delta t=1/640. Fig. 2 and 3 show the evolution of surface fraction and vorticity for three different values of line tension: σγ=0,0.01,0.1\sigma_{\gamma}=0,0.01,0.1. The evolution of both quantities does not vary significantly when going from σγ=0\sigma_{\gamma}=0 to σγ=0.01\sigma_{\gamma}=0.01, although some differences can be noticed from t=4.5t=4.5 on. Changing to σγ=0.1\sigma_{\gamma}=0.1 produces more evident differences, starting already from t=1.375t=1.375. When σγ0\sigma_{\gamma}\neq 0, the NS part of the CHNS system is two-way coupled to the CH part. So, the differences are significant both for surface fraction and vorticity.

\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch0-t1_cut.png} \put(-95.0,45.0){$\sigma_{\gamma}=0$} \put(30.0,110.0){$t=1$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch0-t1375_cut-min.png} \put(0.0,110.0){$t=1.375$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch0-t2_cut-min.png} \put(30.0,110.0){$t=2$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch0-t45_cut-min.png} \put(15.0,110.0){$t=4.5$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch0-t55_cut-min.png} \put(15.0,110.0){$t=5.5$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch0-t85_cut-min.png} \put(15.0,110.0){$t=8.5$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch0-t10_cut-min.png} \put(15.0,110.0){$t=10$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch001-t1_cut-min.png} \put(-95.0,45.0){$\sigma_{\gamma}=0.01$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch001-t1375_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch001-t2_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch001-t45_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch001-t55_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch001-t85_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch001-t10_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch01-t1_cut-min.png} \put(-95.0,45.0){$\sigma_{\gamma}=0.1$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch01-t1375_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch01-t2_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch01-t45_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch01-t55_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch01-t85_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khch01-t10_cut-min.png} \end{overpic}
\begin{overpic}[width=216.81pt,grid=false,tics=10]{images/rgb.png} \end{overpic}
Figure 2: KH instability: evolution of order parameter for different values of line tension: σ=0\sigma=0 (top), σ=0.01\sigma=0.01 (center), and σ=0.1\sigma=0.1 (bottom). A full animation can be viewed following the link youtu.be/C33_{-}WLO1Wd7Y
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh0-t1_cut.png} \put(-95.0,45.0){$\sigma_{\gamma}=0$} \put(28.0,105.0){$t=1$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh0-t1375_cut-min.png} \put(0.0,105.0){$t=1.375$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh0-t2_cut-min.png} \put(28.0,105.0){$t=2$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh0-t45_cut-min.png} \put(18.0,105.0){$t=4.5$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh0-t55_cut-min.png} \put(18.0,105.0){$t=5.5$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh0-t85_cut-min.png} \put(18.0,105.0){$t=8.5$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh0-t10_cut-min.png} \put(23.0,105.0){$t=10$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh001-t1_cut-min.png} \put(-95.0,45.0){$\sigma_{\gamma}=0.01$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh001-t1375_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh001-t2_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh001-t45_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh001-t55_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh001-t85_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh001-t10_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh01-t1_cut-min.png} \put(-95.0,45.0){$\sigma_{\gamma}=0.1$} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh01-t1375_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh01-t2_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh01-t45_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh01-t55_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh01-t85_cut-min.png} \end{overpic}
\begin{overpic}[width=47.69846pt,grid=false,tics=10]{images/KH/khwh01-t10_cut-min.png} \end{overpic}
\begin{overpic}[width=216.81pt,grid=false,tics=10]{images/hsv.png} \end{overpic}
Figure 3: KH instability: evolution of the vorticity for different values of line tension: σγ=0\sigma_{\gamma}=0 (top), σγ=0.01\sigma_{\gamma}=0.01 (center), and σγ=0.1\sigma_{\gamma}=0.1 (bottom). A full animation can be viewed following the link youtu.be/FdMznBuMJPE

5.3.   The Rayleigh–Taylor instability

The Rayleigh–Taylor (RT) instability occurs when a gravity force is acting on a heavier fluid that lies above a lighter fluid. As the RT instability develops, “plumes” of the lighter fluid flow upwards (with respect to the gravitational field) and “spikes” of the heavier fluid fall downwards. We will simulate the RT instability on a sphere and on a torus with the aim of investigating the effect of the geometry. In addition, we will vary line tension and fluid viscosity.

We take two fluids with densities ρ2=3\rho_{2}=3, ρ1=1\rho_{1}=1 and matching viscosities η1=η2=η\eta_{1}=\eta_{2}=\eta, which will be specified for each test. The initial surface fraction is given by

c0=12(1+tanhz+zrand22ϵ),c_{0}=\frac{1}{2}\left(1+\tanh\frac{z+z_{rand}}{2\sqrt{2}\epsilon}\right),

where ϵ=0.025\epsilon=0.025 and zrandz_{rand} is a uniformly generated random number from the range (0.1ϵ,0.1ϵ)(-0.1\epsilon,0.1\epsilon). The role of the perturbation generated by zrandz_{rand} is to onset the RT instability. We set M=0.0025M=0.0025.

Let us start with the sphere. We select mesh level =5\ell=5 (see mesh description for the convergence test) and set Δt=0.1\Delta t=0.1. Fig. 5 shows the evolution of the surface fractions and velocity field for η=102\eta=10^{-2} and two values of line tension: σγ=0\sigma_{\gamma}=0 and σγ=0.025\sigma_{\gamma}=0.025. At time t=7t=7, for σγ=0.025\sigma_{\gamma}=0.025 we observe the characteristic flow structures of the RT instability. Instead, for σγ=0\sigma_{\gamma}=0 such structures have already broken up at t=7t=7. The effect of line tension is also seen at t=30t=30: for σγ=0.025\sigma_{\gamma}=0.025 we observe that the heavier fluid has already settled at the bottom of the sphere, while for σγ=0\sigma_{\gamma}=0 that has not happened yet. It takes till t=55t=55 to have the heavier fluid at the bottom in the absence of line tension. After the revolution, the fluid phases do not achieve steady state quickly but the waves keep traveling along the equator.

\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_sphere/t00_sigma=0_cropped-min.png} \put(-85.0,45.0){$\sigma_{\gamma}=0$} \put(30.0,105.0){$t=0$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_sphere/t07_sigma=0_cropped-min.png} \put(30.0,105.0){$t=7$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_sphere/t14_sigma=0_cropped-min.png} \put(27.0,105.0){$t=14$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_sphere/t20_sigma=0_cropped-min.png} \put(27.0,105.0){$t=20$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_sphere/t30_sigma=0_cropped-min.png} \put(27.0,105.0){$t=30$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_sphere/t55_sigma=0_cropped-min.png} \put(27.0,105.0){$t=55$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_sphere/t00_sigma=0025_cropped-min.png} \put(-85.0,45.0){$\sigma_{\gamma}=0.025$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_sphere/t07_sigma=0025_cropped-min.png} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_sphere/t14_sigma=0025_cropped-min.png} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_sphere/t20_sigma=0025_cropped-min.png} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_sphere/t30_sigma=0025_cropped-min.png} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_sphere/t55_sigma=0025_cropped-min.png} \end{overpic}
\begin{overpic}[width=216.81pt,grid=false,tics=10]{images/lab.png} \end{overpic}
Figure 4: RT instability on the sphere: evolution of the order parameter (color) and velocity field (arrows) for η=102\eta=10^{-2} and different values of line tension: σγ=0\sigma_{\gamma}=0 (top) and σγ=0.025\sigma_{\gamma}=0.025 (bottom). A full animation can be viewed following the link youtu.be/-OmR--qxvAI

Next, we consider an assymetric torus with constant distant from the center of the tube to the origin R=1R=1 and variable radius of the tube: rmin=0.3r(x,y)rmax=0.6r_{min}=0.3\leq r(x,y)\leq r_{max}=0.6, with r(x,y)=rmin+0.5(rmaxrmin)(1xx2+y2)r(x,y)=r_{min}+0.5(r_{max}-r_{min})(1-\frac{x}{\sqrt{x^{2}+y^{2}}}). We characterize the torus surface as the zero level set of function ϕ=(x2+y2+z2+R2r(x,y)2)24R2(x2+y2)\phi=(x^{2}+y^{2}+z^{2}+R^{2}-r(x,y)^{2})^{2}-4R^{2}(x^{2}+y^{2}). The torus is embedded in an outer domain Ω=[5/3,5/3]3\Omega=[-5/3,5/3]^{3}, just like the sphere. We also selected same mesh level (l=5l=5) and same time step (Δt=0.1\Delta t=0.1) as for the sphere. We set the line tension to σ=0.025\sigma=0.025 and vary the viscosity: η=102,101,1\eta=10^{-2},10^{-1},1. Fig. 5 displays the evolution of the surface fractions for these three values of viscosity. First, we observe that in all cases the instability develops more slowly on the “skinny” side of the torus. See second column in Fig. 5. The fact that geometry has a considerable effect on the surface RT instability is also clear when one compares the results on the sphere and the torus for the same values of σγ\sigma_{\gamma} and η\eta, i.e. the top row in Fig. 5 with the bottom row in Fig. 4. In particular, notice that while the heavier fluid reaches the bottom of the sphere around t=30t=30 (Fig. 4, bottom second-last panel), the two fluids are still very much mixed on the torus at t=160t=160 (Fig. 5, top left panel). We need to increase the viscosity value to 1 to be able to see most of the heavier fluid at the bottom of the torus at t=160t=160 (Fig. 5, bottom left panel), although that is still far from being settled.

\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=001_t000-min.png} \put(-70.0,45.0){$\eta=10^{-2}$} \put(30.0,100.0){$t=0$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=001_t002-min.png} \put(30.0,100.0){$t=2$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=001_t004-min.png} \put(30.0,100.0){$t=4$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=001_t010-min.png} \put(27.0,100.0){$t=10$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=001_t040-min.png} \put(27.0,100.0){$t=40$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=001_t160-min.png} \put(25.0,100.0){$t=160$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=01_t000-min.png} \put(-70.0,45.0){$\eta=10^{-1}$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=01_t002-min.png} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=01_t004-min.png} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=01_t010-min.png} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=01_t040-min.png} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=01_t160-min.png} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=1_t000-min.png} \put(-70.0,45.0){$\eta=1$} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=1_t002-min.png} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=1_t004-min.png} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=1_t010-min.png} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=1_t040-min.png} \end{overpic}
\begin{overpic}[width=60.70653pt,grid=false,tics=10]{images/rt_torus/rt_lt=1_nu=1_t160-min.png} \end{overpic}
\begin{overpic}[width=216.81pt,grid=false,tics=10]{images/lab.png} \end{overpic}
Figure 5: RT instability on the torus: evolution of the order parameter for σγ=0.025\sigma_{\gamma}=0.025 and different values of viscosity: η=102\eta=10^{-2} (top), η=101\eta=10^{-1} (center), and η=1\eta=1 (bottom). A full animation can be viewed following the link youtu.be/FTqqFjvzEZg

6 Conclusions

In this paper, we presented an extension of a well-known phase field model for two-phase incompressible flow, and we applied and analyzed an unfitted finite element method for its numerical approximation. The advantage of our model is that it is thermodynamically consistent for a general monotone relation of density and phase-field variable. Because of our interest in biomembranes, this Navier–Stokes–Cahn–Hilliard type system is posed on an arbitrary-shaped closed smooth surface. Although in this paper we assumed the surfaces to be rigid, our long term goal is to study two-phase incompressible flow on evolving shapes. In fact, biological membranes exhibit shape changes that need to be accounted for in a realistic model. This need dictated our choice for the numerical approach, which is a versatile geometrically unfitted finite element method called TraceFEM.

In order to reduce the computational cost, the discrete scheme we proposed decouples the fluid problem (a linearized Navier–Stokes type problem) from the phase-field problem (a Cahn–Hilliard type problem with constant mobility) at each time step. An attractive feature of our scheme is that the numerical solution satisfies the same stability bound as the solution of the original system under some restrictions on the discretization parameters.

We validated our implementation of the proposed numerical scheme with a benchmark problem and applied it to simulate well-known two-phase fluid flows: the Kelvin–Helmholtz and Rayleigh–Taylor instabilities. We investigated the effect of line tension on such instabilities. For the Rayleigh–Taylor instability, we also assessed the effect of viscosity and surface shape, which plays an important role in the evolution of the instability.

Acknowledgments

This work was partially supported by US National Science Foundation (NSF) through grant DMS-1953535. A.Z. and M.O.  also acknowledge the support from NSF through DMS-2011444. A.Q. also acknowledges the support from NSF through DMS-1620384.

References

  • [1] H. Abels and D. Breit. Weak solutions for a non-newtonian diffuse interface model with different densities. Nonlinearity, 29(11):3426, 2016.
  • [2] H. Abels, H. Garcke, and G. Grün. Thermodynamically consistent, frame indifferent diffuse interface models for incompressible two-phase flows with different densities. Mathematical Models and Methods in Applied Sciences, 22(03):1150013, 2012.
  • [3] H. Abels, H. Garcke, and J. Weber. Existence of weak solutions for a diffuse interface model for two-phase flow with surfactants. Communications on Pure & Applied Analysis, 18(1):195–225, 2019.
  • [4] G. L. Aki, W. Dreyer, J. Giesselmann, and C. Kraus. A quasi-incompressible diffuse interface model with phase transition. Mathematical Models and Methods in Applied Sciences, 24(05):827–861, 2014.
  • [5] A. Bonito, A. Demlow, and M. Licht. A divergence-conforming finite element method for the surface stokes equation. SIAM Journal on Numerical Analysis, 58(5):2764–2798, 2020.
  • [6] F. Boyer. A theoretical and numerical model for the study of incompressible mixture flows. Computers & Fluids, 31(1):41–68, 2002.
  • [7] H. Ding, P. D. Spelt, and C. Shu. Diffuse interface model for incompressible two-phase flows with large density ratios. Journal of Computational Physics, 226(2):2078–2095, 2007.
  • [8] S. Dong and J. Shen. A time-stepping scheme involving constant coefficient matrices for phase-field simulations of two-phase incompressible flows with large density ratios. Journal of Computational Physics, 231(17):5788–5804, 2012.
  • [9] Q. Du, L. Ju, and L. Tian. Finite element approximation of the Cahn–Hilliard equation on surfaces. Computer Methods in Applied Mechanics and Engineering, 200(29-32):2458–2470, 2011.
  • [10] C. Eilks and C. Elliott. Numerical simulation of dealloying by surface dissolution via the evolving surface finite element method. Journal of Computational Physics, 227(23):9727 – 9741, 2008.
  • [11] H. Emmerich. The Diffuse Interface Approach in Materials Science: Thermodynamic Concepts and Applications of Phase-Field Models. Springer Publishing Company, Incorporated, 2011.
  • [12] X. Feng. Fully discrete finite element approximations of the navier–stokes–cahn-hilliard diffuse interface model for two-phase fluid flows. SIAM Journal on Numerical Analysis, 44(3):1049–1072, 2006.
  • [13] H. Garcke, M. Hinze, and C. Kahle. A stable and linear time discretization for a thermodynamically consistent model for two-phase incompressible flow. Applied Numerical Mathematics, 99:151–171, 2016.
  • [14] Y. Gong, J. Zhao, and Q. Wang. An energy stable algorithm for a quasi-incompressible hydrodynamic phase-field model of viscous fluid mixtures with variable densities and viscosities. Computer Physics Communications, 219:20–34, 2017.
  • [15] Y. Gong, J. Zhao, X. Yang, and Q. Wang. Fully discrete second-order linear schemes for hydrodynamic phase field models of binary viscous fluid flows with variable densities. SIAM Journal on Scientific Computing, 40(1):B138–B167, 2018.
  • [16] J. Grande, C. Lehrenfeld, and A. Reusken. Analysis of a high-order trace finite element method for PDEs on level set surfaces. SIAM Journal on Numerical Analysis, 56(1):228–255, 2018.
  • [17] G. Grun. On convergent schemes for diffuse interface models for two-phase flow of incompressible fluids with general mass densities. SIAM Journal on Numerical Analysis, 51(6):3036–3061, 2013.
  • [18] J.-L. Guermond and L. Quartapelle. A projection fem for variable density incompressible flows. Journal of Computational Physics, 165(1):167–188, 2000.
  • [19] Z. Guo, P. Lin, J. Lowengrub, and S. Wise. Mass conservative and energy stable finite difference methods for the quasi-incompressible Navier–Stokes–Cahn–Hilliard system: Primitive variable and projection-type schemes. Computer Methods in Applied Mechanics and Engineering, 326:144–174, 2017.
  • [20] M. E. Gurtin and A. I. Murdoch. A continuum theory of elastic material surfaces. Archive for Rational Mechanics and Analysis, 57(4):291–323, 1975.
  • [21] D. Han, A. Brylev, X. Yang, and Z. Tan. Numerical analysis of second order, fully discrete energy stable schemes for phase field models of two-phase incompressible flows. Journal of Scientific Computing, 70(3):965–989, 2017.
  • [22] P. C. Hohenberg and B. I. Halperin. Theory of dynamic critical phenomena. Rev. Mod. Phys., 49:435–479, Jul 1977.
  • [23] T. Jankuhn, M. A. Olshanskii, and A. Reusken. Incompressible fluid problems on embedded surfaces: Modeling and variational formulations. Interfaces and Free Boundaries, 20(3):353–377, 2018.
  • [24] T. Jankuhn, M. A. Olshanskii, A. Reusken, and A. Zhiliakov. Error analysis of higher order trace finite element methods for the surface Stokes equation. Journal of Numerical Mathematics, 2020.
  • [25] D. Kay, V. Styles, and R. Welford. Finite element approximation of a cahn–hilliard–navier–stokes system. Interfaces and Free Boundaries, 10(1):15–43, 2008.
  • [26] P. L. Lederer, C. Lehrenfeld, and J. Schöberl. Divergence-free tangential finite element methods for incompressible flows on surfaces. International Journal for Numerical Methods in Engineering, 121(11):2503–2533, 2020.
  • [27] C. Lehrenfeld, M. A. Olshanskii, and X. Xu. A stabilized trace finite element method for partial differential equations on evolving surfaces. SIAM Journal on Numerical Analysis, 56(3):1643–1672, 2018.
  • [28] C. Liu, F. Frank, C. Thiele, F. O. Alpak, S. Berg, W. Chapman, and B. Riviere. An efficient numerical algorithm for solving viscosity contrast Cahn-–Hilliard–-Navier-–Stokes system in porous media. Journal of Computational Physics, 400:108948, 2020.
  • [29] J. Lowengrub and L. Truskinovsky. Quasi–incompressible Cahn–Hilliard fluids and topological transitions. Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, 454(1978):2617–2654, 1998.
  • [30] I. Nitschke, A. Voigt, and J. Wensch. A finite element approach to incompressible two-phase flow on manifolds. Journal of Fluid Mechanics, 708:418–438, 2012.
  • [31] R. H. Nochetto, A. J. Salgado, and I. Tomas. A diffuse interface model for two-phase ferrofluid flows. Computer Methods in Applied Mechanics and Engineering, 309:497–531, 2016.
  • [32] M. Olshanskii, A. Quaini, A. Reusken, and V. Yushutin. A finite element method for the surface stokes problem. SIAM Journal on Scientific Computing, 40(4):A2492–A2518, 2018.
  • [33] M. Olshanskii, A. Reusken, and A. Zhiliakov. Inf-sup stability of the trace P2-P1 Taylor–Hood elements for surface PDEs. Mathematics of Computation, 2021.
  • [34] M. Olshanskii, A. Reusken, and A. Zhilikov. Recycling augmented Lagrangian preconditioner in an incompressible fluid solver. arXiv:2012.10073, 2020.
  • [35] M. A. Olshanskii. A low order Galerkin finite element method for the Navier–Stokes equations of steady incompressible flow: a stabilization issue and iterative methods. Computer Methods in Applied Mechanics and Engineering, 191(47-48):5515–5536, 2002.
  • [36] M. A. Olshanskii, A. Reusken, and J. Grande. A finite element method for elliptic equations on surfaces. SIAM Journal on Numerical Analysis, 47:3339–3358, 2009.
  • [37] A. Reusken. Analysis of trace finite element methods for surface partial differential equations. IMA Journal of Numerical Analysis, 35(4):1568–1590, 2015.
  • [38] J. Shen, J. Xu, and J. Yang. The scalar auxiliary variable (sav) approach for gradient flows. Journal of Computational Physics, 353:407–416, 2018.
  • [39] J. Shen and X. Yang. Numerical approximations of Allen-Cahn and Cahn-Hilliard equations. Discrete & Continuous Dynamical Systems - A, 28:1669, 2010.
  • [40] J. Shen and X. Yang. A phase-field model and its numerical approximation for two-phase incompressible flows with different densities and viscosities. SIAM Journal on Scientific Computing, 32(3):1159–1179, 2010.
  • [41] J. Shen and X. Yang. Decoupled, energy stable schemes for phase-field models of two-phase incompressible flows. SIAM Journal on Numerical Analysis, 53(1):279–296, 2015.
  • [42] M. Shokrpour Roudbari, G. Şimşek, E. H. van Brummelen, and K. G. van der Zee. Diffuse-interface two-phase flow models with different densities: A new quasi-incompressible form and a linear energy-stable method. Mathematical Models and Methods in Applied Sciences, 28(04):733–770, 2018.
  • [43] J. Yang and J. Kim. A phase-field model and its efficient numerical method for two-phase flows on arbitrarily curved surfaces in 3d space. Computer Methods in Applied Mechanics and Engineering, 372:113382, 2020.
  • [44] J. Yang, S. Mao, X. He, X. Yang, and Y. He. A diffuse interface model and semi-implicit energy stable finite element method for two-phase magnetohydrodynamic flows. Computer Methods in Applied Mechanics and Engineering, 356:435–464, 2019.
  • [45] V. Yushutin, A. Quaini, S. Majd, and M. Olshanskii. A computational study of lateral phase separation in biological membranes. International journal for numerical methods in biomedical engineering, 35(3):e3181, 2019.
  • [46] A. Zhiliakov, Y. Wang, A. Quaini, M. Olshanskii, and S. Majd. Experimental validation of a phase-field model to predict coarsening dynamics of lipid domains in multicomponent membranes. Biochimica et Biophysica Acta (BBA)-Biomembranes, 1863(1):183446, 2021.
  • [47] G. Zhu, H. Chen, A. Li, S. Sun, and J. Yao. Fully discrete energy stable scheme for a phase-field moving contact line model with variable densities and viscosities. Applied Mathematical Modelling, 83:614–639, 2020.
  • [48] G. Zhu, H. Chen, J. Yao, and S. Sun. Efficient energy-stable schemes for the hydrodynamics coupled phase-field model. Applied Mathematical Modelling, 70:82–108, 2019.