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

A Uniquely Solvable, Positivity-Preserving and Unconditionally Energy Stable Numerical Scheme for the Functionalized Cahn-Hilliard Equation with Logarithmic Potential

Wenbin Chen111 School of Mathematical Sciences and Shanghai Key Laboratory for Contemporary Applied Mathematics, Fudan University, Shanghai 200433, China [email protected] Jianyu Jing222 School of Mathematical Sciences, Fudan University, Shanghai 200433, China [email protected] Hao Wu333 School of Mathematical Sciences and Shanghai Key Laboratory for Contemporary Applied Mathematics, Fudan University, Shanghai 200433, China [email protected]
Abstract

We propose and analyze a first-order finite difference scheme for the functionalized Cahn-Hilliard (FCH) equation with a logarithmic Flory-Huggins potential. The semi-implicit numerical scheme is designed based on a suitable convex-concave decomposition of the FCH free energy. We prove unique solvability of the numerical algorithm and verify its unconditional energy stability without any restriction on the time step size. Thanks to the singular nature of the logarithmic part in the Flory-Huggins potential near the pure states ±1\pm 1, we establish the so-called positivity-preserving property for the phase function at a theoretic level. As a consequence, the numerical solutions will never reach the singular values ±1\pm 1 in the point-wise sense and the fully discrete scheme is well defined at each time step. Next, we present a detailed optimal rate convergence analysis and derive error estimates in l(0,T;Lh2)l2(0,T;Hh3)l^{\infty}(0,T;L_{h}^{2})\cap l^{2}(0,T;H^{3}_{h}) under a linear refinement requirement ΔtC1h\Delta t\leq C_{1}h. To achieve the goal, a higher order asymptotic expansion (up to the second order temporal and spatial accuracy) based on the Fourier projection is utilized to control the discrete maximum norm of solutions to the numerical scheme. We show that if the exact solution to the continuous problem is strictly separated from the pure states ±1\pm 1, then the numerical solutions can be kept away from ±1\pm 1 by a positive distance that is uniform with respect to the size of the time step and the grid. Finally, a few numerical experiments are presented. Convergence test is performed to demonstrate the accuracy and robustness of the proposed numerical scheme. Pearling bifurcation, meandering instability and spinodal decomposition are observed in the numerical simulations.

keywords:
Functionalized Cahn-Hilliard equation, Finite difference scheme, Logarithmic potential, Unique solvability, Energy stability, Positivity preserving, Optimal rate convergence analysis, higher order asymptotic expansion
MSC:
[2020] 35K35 , 35K55 , 65M06 , 65M12
journal: XXX

1 Introduction

We consider the following functionalized Cahn-Hilliard (FCH) equation:

{tϕ=Δμ,μ=ϵ2Δω+f(ϕ)ωϵpηω,ω=ϵ2Δϕ+f(ϕ),inΩ×(0,T).\begin{cases}&\partial_{t}\phi=\Delta\mu,\\ &\mu=-\epsilon^{2}\Delta\omega+f^{\prime}(\phi)\omega-\epsilon^{p}\eta\omega,\\ &\omega=-\epsilon^{2}\Delta\phi+f(\phi),\end{cases}\qquad\text{in}\ \Omega\times(0,T). (1.1)

Here, we assume that T>0T>0 is a given final time and Ω=(0,1)d\Omega=(0,1)^{d} with the spatial dimension 1d31\leq d\leq 3. For the sake of simplicity, below we analyze the system (1.1) in a periodic setting on Ω\Omega such that the variables ϕ\phi, μ\mu, ω\omega satisfy periodic boundary conditions. With some minor modifications, our results can be extended to the case with homogeneous Neumann boundary conditions, or boundary conditions of mixed periodic-homogeneous Neumann type.

In the system (1.1), the scalar function ϕ:Ω×(0,T)\phi:\Omega\times(0,T)\to\mathbb{R} denotes local concentrations of the two components in a binary mixture (e.g., an amphiphilic material in a solvent). Here we define ϕ\phi as the difference of volume fractions. In view of its physical interpretation, only the values ϕ[1,1]\phi\in[-1,1] are admissible. The pure states correspond to the values 1-1 (the pure solvent) and 11 (the pure amphiphile), whereas ϕ(1,1)\phi\in(-1,1) indicates the presence of a mixing state. The scalar function μ:Ω×(0,T)\mu:\Omega\times(0,T)\to\mathbb{R} is referred to as the chemical potential, which is the first variational derivative of the following free energy functional:

E(ϕ)=Ω12(ϵ2Δϕf(ϕ))2ϵpη(ϵ22|ϕ|2+F(ϕ))dx,E(\phi)=\int_{\Omega}\frac{1}{2}\left(\epsilon^{2}\Delta\phi-f(\phi)\right)^{2}-\epsilon^{p}\eta\left(\frac{\epsilon^{2}}{2}|\nabla\phi|^{2}+F(\phi)\right)\,\mathrm{d}x, (1.2)

where ϵ>0\epsilon>0, η>0\eta>0 and p{1,2}p\in\{1,2\} are given parameters. The energy (1.2) extends the classical Cahn-Hilliard energy in the literature (see [1])

ECH(ϕ)=Ωϵ22|ϕ|2+F(ϕ)dx,E_{\mathrm{CH}}(\phi)=\int_{\Omega}\frac{\epsilon^{2}}{2}|\nabla\phi|^{2}+F(\phi)\,\mathrm{d}x, (1.3)

where FF stands for the (Helmholtz) free energy density of mixing and its derivative is denoted by f=Ff=F^{\prime}. Then the scalar function ω:Ω×(0,T)\omega:\Omega\times(0,T)\to\mathbb{R} in (1.1) corresponds to the first variational derivative of ECHE_{\mathrm{CH}}, which is the chemical potential in the classical Cahn-Hilliard theory for binary mixtures.

In (1.2), the positive parameter ϵ1\epsilon\ll 1 characterizes the width of inner structures, e.g., the (diffuse) interfaces between two components. Next, let us pay some more attention on the parameter η\eta\in\mathbb{R}, which plays an important role in the modelling and analysis. Indeed, the sign of η\eta can change fundamentally the structure of the problem as well as the physically motivating examples [2]. When η=0\eta=0, the energy EE reduces to the well-known phase-field formulation of the Willmore functional (PFW) that approximates the Canham-Helfrich bending energy of surfaces. The corresponding phase-field model has been used to study deformations of elastic vesicles subject to possible volume/surface constraints (see e.g., [3, 4]). When η<0\eta<0, the energy EE is related to the so-called Cahn-Hilliard-Willmore (CHW) energy, which was introduced in [5, 6, 7] to investigate strong anisotropy effects arising in the growth and coarsening of thin films. In this paper, our interest will focus on the case η>0\eta>0 such that EE is referred to as the functionalized Cahn-Hilliard (FCH) energy in the literature. It was first proposed in [8] to model phase separation of mixtures with an amphiphilic structure. Later on, the extended FCH model was used to describe nanoscale morphology changes in functionalized polymer chains [9], amphiphilic bilayer interfaces [2, 10, 11], and membranes undergoing pearling bifurcations [12, 13]. With a positive parameter η\eta, the FCH energy (1.2) reflects the balance between the square of the variational derivative of ECHE_{\mathrm{CH}} against itself: the first term stabilizes bilayers, pores and micelle structures, while the second term promotes the growth of interfaces and competition between these geometries (see e.g., [14]). Minimizers of the FCH energy are approximate critical points of ECHE_{\mathrm{CH}}, which however favor more surface area. This fact dramatically changes the nature of the energy landscape and presents rich morphological structures that are quite different from the phase separation process described by the classical Cahn-Hilliard energy. The FCH energy can naturally incorporate the propensity of the amphiphilic surfactant phase to drive the creation of stable bilayers, or homoclinic interfaces with an intrinsic width [2]. In particular, the bilayer structure has the distinct feature that it separate two identical phases by a thin region of a second phase. For recent works on the minimization problem of the FCH energy, bilayer-structures, pearled patterns, nanostructure morphology changes and network bifurcations, we refer to [13, 14, 15, 16, 17] and the references cited therein. Finally, we note that the exponent pp in (1.2) distinguishes some natural limits of the FCH energy, when p=1p=1, it represents the strong functionalization, while the choice p=2p=2 corresponds to the weak functionalization (see [10, 12, 15] for detailed discussions).

In this study, we work with the following thermodynamically relevant potential of logarithmic type:

F(r)=(1+r)ln(1+r)+(1r)ln(1r)λ2r2,r(1,1),F(r)=(1+r)\ln(1+r)+(1-r)\ln(1-r)-\frac{\lambda}{2}r^{2},\qquad r\in(-1,1), (1.4)

where λ\lambda\in\mathbb{R} is a given parameter. The potential function FF was proposed in the original work [1] on the phase separation of binary alloys and it also arises in the Flory–Huggins theory for polymer solutions [18]. The first order derivative of FF is denoted by ff such that

f(r)=ln1+r1rλr,r(1,1).f(r)=\ln\frac{1+r}{1-r}-\lambda r,\qquad r\in(-1,1). (1.5)

When λ>2\lambda>2, FF has a double-well structure with two minima ±r(1,1)\pm\,r_{*}\in(-1,1), where rr_{*} is the positive root of the equation f(r)=0f(r)=0. Besides, we observe that limr±1f(r)=±\lim_{r\to\pm 1}f(r)=\pm\infty. This singular nature of ff near the pure states ±1\pm 1 brings great challenges in the study of related phase-field equations, including the classical Cahn-Hilliard equation driven by ECHE_{\mathrm{CH}}. On the other hand, to avoid possible difficulties caused by the singularity of ff, a widely used approach is to work with an approximate polynomial like

F(r)=14(r21)2,r,F(r)=\frac{1}{4}(r^{2}-1)^{2},\quad r\in\mathbb{R}, (1.6)

which is sometimes called the “shallow quenching” approximation of the logarithmic potential (1.4) (see e.g., [19, 20]). Concerning the Cahn-Hilliard equation tϕ=[(ϵ2Δϕ+f(ϕ))]\partial_{t}\phi=\nabla\cdot[\mathcal{M}\nabla(-\epsilon^{2}\Delta\phi+f(\phi))] with logarithmic potential (1.4), we refer to [19, 21, 22, 23, 24, 25, 26] for extensive theoretic analysis on well-posedness as well as long-time behavior of global solutions (see also the review articles [20, 27]), while for contributions in the numerical studies, we recall [28, 29, 30, 31, 32, 33, 34, 35, 36] and the references therein.

Under suitable boundary conditions (e.g., the periodic or Neumann type), the FCH equation (1.1) can be viewed as an H1H^{-1} gradient flow of the FCH energy (1.2), which not only dissipates the energy but also preserves the mass along time evolution. As long as a solution ϕ\phi exists on (0,T)(0,T) and is sufficiently regular, it holds

ddtE(ϕ)=Ω|μ|2dx0,t(0,T),\frac{\textrm{d}}{\textrm{d}t}E(\phi)=-\int_{\Omega}\left|\nabla\mu\right|^{2}\textrm{d}x\leq 0,\quad\forall\,t\in(0,T), (1.7)

and

Ωϕ(,t)dx=Ωϕ(,0)dx,t[0,T].\displaystyle\int_{\Omega}\phi(\cdot,t)\,\mathrm{d}x=\int_{\Omega}\phi(\cdot,0)\,\mathrm{d}x,\quad\forall\,t\in[0,T]. (1.8)

Concerning the theoretic analysis of equation (1.1), when η>0\eta>0 and the periodic boundary conditions are imposed, Dai et al. [37] proved the existence of global weak solutions in the case with a general regular potential including (1.6) and a degenerate mobility =(ϕ)\mathcal{M}=\mathcal{M}(\phi) (see also [38]). For the case with the regular potential (1.6) and a constant mobility, existence and uniqueness of global solutions in the Gevrey class were established in [39] for η\eta\in\mathbb{R}, again in the periodic setting. We also mention [40, 41] in which some theoretic results for the Cahn-Hilliard-Willmore equation (with η<0\eta<0) can be found. The case associated with a logarithmic potential (1.4) is more difficult. In the recent work [42], Schimperna and Wu investigated the initial boundary value problem of (1.1) subject to homogeneous Neumann boundary conditions with η\eta\in\mathbb{R}. After overcoming difficulties from the singular potential and its interaction with higher-order spatial derivatives, they proved existence, uniqueness and regularity of a global weak solution and characterized its long-time behavior. For a higher-order parabolic equation, in general we do not have maximum principle for its solution. Nevertheless, the singular character of the nonlinearity ff can help to keep the weak solutions of (1.1) staying away from the pure states at a point-wise level such that 1<ϕ(x,t)<1-1<\phi(x,t)<1 almost everywhere in Ω×(0,T)\Omega\times(0,T). This is sometimes referred to as a positivity-preserving property, since both 1+ϕ>01+\phi>0 and 1ϕ>01-\phi>0 hold. When the spatial dimension is less than or equal to two, an improved conclusion has been drawn in [42]: if the initial datum ϕ0\phi_{0} is strictly separated from ±1\pm 1, then there exists a uniform distance δ(0,1)\delta\in(0,1) such that ϕ(t)C(Ω¯)1δ\|\phi(t)\|_{C(\overline{\Omega})}\leq 1-\delta for all t[0,T]t\in[0,T]. This fact is crucial for obtaining higher order spatial regularity of solutions, since the singular potential ff is now confined on a compact subset of (1,1)(-1,1) and thus is smooth. We note that similar properties have been proven for the classical Cahn-Hilliard equation with a logarithmic potential and for some extended systems, see for instance, [24, 25, 26] and the references therein.

Based on the analytic results obtained in [42], we are going to provide a detailed numerical analysis of the FCH equation (1.1) with logarithmic potential (1.4), subject to periodic boundary conditions. Observe that (1.1) is a sixth order parabolic equation for ϕ\phi (see (2.3)–(2.4) below). Those nonlinear terms in the chemical potential μ\mu such as β(ϕ)β(ϕ)\beta(\phi)\beta^{\prime}(\phi), β′′(ϕ)|ϕ|2\beta^{\prime\prime}(\phi)|\nabla\phi|^{2}, Δβ(ϕ)\Delta\beta(\phi) with β(ϕ)f(ϕ)+λϕ\beta(\phi)\triangleq f(\phi)+\lambda\phi (see (2.4)) present a highly nonlinear and singular nature of the problem, which turns out to be more involved than the fourth order Cahn-Hilliard equation. This feature yields great challenges in the numerical study. One obvious difficulty is to overcome the step restriction in the time-space discretization such that an explicit numerical scheme will require a rather restrictive Courant-Friedrichs-Lewy (CFL) condition like ΔtCh6\Delta t\leq Ch^{6}, with Δt\Delta t and hh being the time and spatial step sizes, respectively. On the other hand, a fully implicit scheme like backward Euler may be difficult to prove its unique solvability and only conditionally energy stable. Our goal is to present a numerical scheme that is uniquely solvable, preserving the mass conservation, energy dissipation and the positivity property, under some mild refinement requirement. Besides, we shall apply some efficient numerical solvers to solve the numerical scheme in the long-time scale simulation.

There have been some progresses on the numerical study of the FCH equation (1.1) with η>0\eta>0 and a polynomial potential like (1.6). In [14], Jones proposed a semi-implicit numerical scheme and proved its energy stability but the unique solvability was not given. He also made comparisons among semi-implicit, fully-implicit and explicit exponential time differencing (ETD) numerical schemes. A fully implicit scheme with pseudo-spectral approximation in space was introduced in Christlieb et al. [43], where the authors performed numerical tests to show its accuracy and efficiency but did not prove energy stability and solvability. In [44], Guo et al. presented a local discontinuous Galerkin method to overcome the difficulty associated with higher order spatial derivatives. Energy stability was shown for the semi-discrete (time-continuous) scheme therein. Later on, a convex-concave splitting scheme was proposed in Feng et al. [45], where the unique solvability and unconditional energy stability were theoretically verified. Besides, the authors performed the convergence analysis and applied an efficient preconditioned steepest descent algorithm to solve their scheme. Recently, theoretical analysis on the energy stability of a stabilized semi-implicit scheme for the FCH equation was carried out in [46]. We also refer to [47] for an unconditionally energy stable second-order numerical scheme based on the scalar auxiliary variable (SAV) approach, to [48] for a highly accurate, linear and unconditionally energy stable large time-stepping scheme, and to [49] for numerical comparison between those schemes based on the stable SAV approach and the classical BDF method.

It is worth mentioning that all the previous numerical studies mentioned above are associated with a regular (polynomial) potential. To the best of our knowledge, there is no result on the numerical analysis of the FCH equation (1.1) with the logarithmic potential (1.4) in the existing literature. Here we aim to make the first attempt in this direction. The main contributions of this paper are summarized below.

(1) We propose a first-order semi-implicit finite difference numerical scheme (see (2.18)–(2.19)), based on a suitable convex-concave decomposition of the highly nonlinear FCH energy (see Proposition 2.3).

One main difficulty comes from the troublesome term Ωβ(ϕ)Δϕdx-\int_{\Omega}\beta(\phi)\Delta\phi\,\mathrm{d}x in the FCH energy, which is neither convex nor concave (cf. (2.8)). Using integration by parts and the periodic boundary conditions, we can rewrite it as Ωβ(ϕ)|ϕ|2dx\int_{\Omega}\beta^{\prime}(\phi)|\nabla\phi|^{2}\,\mathrm{d}x instead. Thanks to the strict separation property of ϕ\phi (see Proposition 2.2), we are able to show that the new functional (and its discrete analogue) is strictly convex on a suitable closed set (cf. [42, 50] for conclusions in some weaker settings). This observation enables us to derive the desired convex splitting scheme that treats the convex part of EE implicitly and the concave part explicitly. Our argument is different from the case with the regular potential (1.6) studied in [45] such that no auxiliary term is required here to guarantee the necessary convexity.

(2) We prove that the proposed numerical scheme (2.18)–(2.19) is uniquely solvable, without any restriction on the size of the time step and the grid (see Theorem 3.1). Moreover, solutions to this fully discrete scheme preserve the mass at each time step (see Proposition 2.4) and satisfy the positivity-preserving property in the point-wise sense (see (3.1)).

The positivity-preserving property at the discrete level is crucial, since it guarantees the numerical scheme to be unconditionally well-defined. Concerning the Cahn-Hilliard equation with logarithmic potential (1.4), this issue was first tackled in [30] where Copetti and Elliott worked with a backward Euler scheme. Under certain constraint on the time step, they proved the unique solvability and positivity-preserving property of numerical solutions by applying a regularization of the singular potential and then passing to the limit. An alternative approach was recently introduced in Chen et al. [29], based on the convex-concave decomposition technique (see e.g., [51, 52]). They proposed a first order semi-implicit convex splitting scheme as well as a second order accurate scheme involving the implicit backward differential formulation (BDF). Then they overcame the time step restriction and theoretically justified the unique solvability and energy stability of those schemes. In particular, the logarithmic part was treated in an implicit way and its convexity as well as singularity helped to prevent the numerical solution from approaching the singular values ±1\pm 1. Later in Chen et al. [28], a similar idea was applied to show the positivity-preserving property of a second order accurate scheme with a modified Crank-Nicolson approximation. Further applications can be found in recent works on the Cahn-Hilliard equation with a Flory-Huggins-de Gennes energy [53, 54], the Poisson-Nernst-Planck system [55] and the reaction-diffusion system with detailed balance [56] etc.

In this study, we successfully generalize the argument in [29] to the sixth order FCH equation (1.1). Roughly speaking, for any nn\in\mathbb{N}, we look for a discrete solution ϕn+1\phi^{n+1} to the numerical scheme (2.18)–(2.19) by solving a minimization problem form certain discrete energy 𝒥n\mathcal{J}^{n} derived from the convex-concave energy decomposition (see (3.2)). If the numerical solution in the previous time step (denoted by ϕn\phi^{n}) takes all its grid values in (1,1)(-1,1) and its discrete average also belongs to (1,1)(-1,1), we can use the specific form of the logarithmic term β(ϕ)\beta(\phi) and in particular, its singular nature near ±1\pm 1 to show that the solution ϕn+1\phi^{n+1} that is a (global) minimum of 𝒥n\mathcal{J}^{n}, must exist and can only possibly occur at an interior point in the numerical solution variable domain 𝒜h\mathcal{A}_{h} (see (3.4)). Besides, uniqueness of the numerical solution is a direct consequence of the strict convexity of the discrete energy function.

(3) We verify the unconditional energy stability of the proposed semi-implicit numerical scheme (2.18)–(2.19), see Theorem 4.2. The proof is based on the convex-concave decomposition of the discrete FCH energy given in Corollary 2.1. Moreover, the unconditional energy stability enables us to derive a uniform in time Hh2H^{2}_{h}-bound for the numerical solutions (see Corollary 4.1), which plays an important role in the subsequent convergence analysis.

(4) We perform an optimal rate convergence analysis for the numerical scheme (2.18)–(2.19) and establish an error estimate with O(Δt+h2)O(\Delta t+h^{2}) accuracy in l(0,T;Lh2)l2(0,T;Hh3)l^{\infty}(0,T;L_{h}^{2})\cap l^{2}(0,T;H^{3}_{h}), under a linear refinement requirement on the time step size such that ΔtC1h\Delta t\leq C_{1}h (see Theorem 5.3).

To obtain the expected error estimate under the above linear refinement constraint, a crucial step is to construct a higher order asymptotic expansion of the exact solution up to the second order temporal and spatial accuracy (see (5.11)). A similar idea was used in [55] for the optimal rate convergence analysis of the Poisson-Nernst-Planck system. Here in our case, based on the lower order error estimate in the discrete space l(0,T;Hh1)l2(0,T;Hh2)l^{\infty}(0,T;H_{h}^{-1})\cap l^{2}(0,T;H^{2}_{h}) (see (5.35)) we can first derive a rough estimate on the modified numerical error function in l(0,T;Lh2)l^{\infty}(0,T;L_{h}^{2}) (see (5.29)), which combined with the inverse inequality further leads to an error estimate in l(0,T;Lh)l^{\infty}(0,T;L_{h}^{\infty}) (see (5.30)), provided that Δt\Delta t and hh are suitably small and satisfy the linear constraint ΔtC1h\Delta t\leq C_{1}h. Furthermore, this LhL_{h}^{\infty} estimate allows us to show that all the numerical solutions can be kept away from the pure states ±1\pm 1 with a uniform positive distance (see (5.31)). Hence, the singular term β\beta (and its derivatives) is confined on a compact subset of (1,1)(-1,1), becoming regular and uniformly bounded. With this crucial observation, we can proceed to derive the required higher order error estimate in l(0,T;Lh2)l2(0,T;Hh3)l^{\infty}(0,T;L_{h}^{2})\cap l^{2}(0,T;H^{3}_{h}) by using the energy method.

The remaining part of this paper is organized as follows. In Section 2 we present some analytic results on the continuous problem and propose the numerical scheme. In Section 3, we provide detailed proofs on unique solvability of the scheme and the positivity-preserving property of numerical solutions. The unconditional energy stability is then established in Section 4. In Section 5, we perform an optimal rate convergence analysis and derive the error estimates. Numerical experiments are presented in Section 6 to verify the accuracy and efficiency of our scheme. The mass conservation, energy dissipation and positivity-preserving properties are verified numerically. In the final Section 7, we make some concluding remarks.

2 The Numerical Scheme

2.1 Preliminaries: well-posedness of the continuous problem

Let XX be a (real) Banach or Hilbert space with the associated norm denoted by X\|\cdot\|_{X}. The symbol XX^{\prime} stands for the dual space of XX and ,\langle\cdot,\cdot\rangle denotes the duality product between XX^{\prime} and XX. Throughout the paper, we consider the domain Ω=(0,1)d\Omega=(0,1)^{d} with 1d31\leq d\leq 3. For mm\in\mathbb{N}, we set

Cperm(Ω¯)={fCm(d)|fisΩ-periodic}andC̊perm(Ω¯)={fCperm(Ω¯)|Ωfdx=0}.\displaystyle C^{m}_{\mathrm{per}}(\overline{\Omega})=\big{\{}f\in C^{m}(\mathbb{R}^{d})\,|\,f\ \text{is}\ \Omega\text{-periodic}\big{\}}\quad\text{and}\quad\mathring{C}^{m}_{\mathrm{per}}(\overline{\Omega})=\Big{\{}f\in C^{m}_{\mathrm{per}}(\overline{\Omega})\,\Big{|}\,\int_{\Omega}f\textrm{d}x=0\Big{\}}.

Next, we recall some standard notations for Lebesgue and Sobolev spaces in the periodic setting (see, e.g., [39]). For p[1,+)p\in[1,+\infty), we define

Lperp(Ω)={fLlocp(d)|fisΩ-periodic}andL̊perp(Ω)={fLperp(Ω)|Ωfdx=0}.\displaystyle L^{p}_{\mathrm{per}}(\Omega)=\big{\{}f\in L^{p}_{\text{loc}}(\mathbb{R}^{d})\,|\,f\ \text{is}\ \Omega\text{-periodic}\big{\}}\quad\text{and}\quad\mathring{L}^{p}_{\mathrm{per}}(\Omega)=\Big{\{}f\in L^{p}_{\mathrm{per}}(\Omega)\,\Big{|}\,\int_{\Omega}f\,\mathrm{d}x=0\Big{\}}.

For m+m\in\mathbb{Z}^{+} and p[1,+)p\in[1,+\infty), we define

Wperm,p(Ω)={fWlocm,p(d)|fisΩ-periodic},W̊perm,p(Ω)=Wperm,p(Ω)L̊perp(Ω).\displaystyle W^{m,p}_{\mathrm{per}}(\Omega)=\{f\in W^{m,p}_{\text{loc}}(\mathbb{R}^{d})\,|\,f\ \text{is}\ \Omega\text{-periodic}\},\quad\mathring{W}^{m,p}_{\mathrm{per}}(\Omega)=W^{m,p}_{\mathrm{per}}(\Omega)\cap\mathring{L}^{p}_{\mathrm{per}}(\Omega).

When p=2p=2, we simply denote Wperm,p(Ω)W^{m,p}_{\mathrm{per}}(\Omega), W̊perm,p(Ω)\mathring{W}^{m,p}_{\mathrm{per}}(\Omega) by Hperm(Ω)H^{m}_{\mathrm{per}}(\Omega), H̊perm(Ω)\mathring{H}^{m}_{\mathrm{per}}(\Omega), respectively. The standard scalar product of Lper2(Ω)L^{2}_{\mathrm{per}}(\Omega) will be denoted by (,)Ω(\cdot,\cdot)_{\Omega}. For m+m\in\mathbb{Z}^{+}, we also define the dual spaces

Hperm(Ω)=(Hperm(Ω))andH̊perm(Ω)={fHperm(Ω)|f,1=0}.H^{-m}_{\mathrm{per}}(\Omega)=(H^{m}_{\mathrm{per}}(\Omega))^{\prime}\quad\text{and}\quad\mathring{H}^{-m}_{\mathrm{per}}(\Omega)=\{f\in H^{-m}_{\mathrm{per}}(\Omega)\,|\,\langle f,1\rangle=0\}.

For the sake of convenience, we denote by BB the convex part of the logarithmic potential FF and by β\beta the monotone part of its derivative ff (see [42]), that is,

B(r):=F(r)+λ2r2=(1+r)ln(1+r)+(1r)ln(1r),r(1,1),\displaystyle B(r):=F(r)+\frac{\lambda}{2}r^{2}=(1+r)\ln(1+r)+(1-r)\ln(1-r),\quad r\in(-1,1), (2.1)
β(r):=f(r)+λr=ln1+r1r,r(1,1).\displaystyle\beta(r):=f(r)+\lambda r=\displaystyle{\ln\frac{1+r}{1-r}},\quad r\in(-1,1). (2.2)

In [42], Schimperna and Wu investigated the system (1.1) in a bounded smooth domain Ωd\Omega\subset\mathbb{R}^{d}, 1d31\leq d\leq 3, subject to the homogeneous Neumann boundary conditions 𝐧ϕ=𝐧ω=𝐧μ=0\partial_{\mathbf{n}}\phi=\partial_{\mathbf{n}}\omega=\partial_{\mathbf{n}}\mu=0 on Ω\partial\Omega (𝐧\mathbf{n} denotes the unit outer normal vector on the boundary Ω\partial\Omega). They proved existence and uniqueness of a global weak solution to the resulting initial boundary value problem and established its regularity for any positive time, see [42, Theorems 2.3, 2.4]. With minor modifications on the proofs therein, we are able to obtain similar results for the FCH equation (1.1) in the periodic setting.

Proposition 2.1 (Well-posedness).

Let Ω=(0,1)d\Omega=(0,1)^{d} with 1d31\leq d\leq 3. Suppose that FF is determined by (1.4), λ,η\lambda,\,\eta\in\mathbb{R} and ϵ>0\epsilon>0 are given constants. For any initial datum ϕ0\phi_{0} satisfying

ϕ0Hper2(Ω),β(ϕ0)Lper2(Ω),Ωϕ0dx(1,1),\phi_{0}\in H^{2}_{\mathrm{per}}(\Omega),\quad\beta(\phi_{0})\in L^{2}_{\mathrm{per}}(\Omega),\quad\int_{\Omega}\phi_{0}\,\mathrm{d}x\in(-1,1),

there exists a unique global weak solution ϕ\phi to the system (1.1) subject to the periodic boundary conditions such that, for any T>0T>0,

ϕH1(0,T;Hper1(Ω))L(0,T;Hper2(Ω))L4(0,T;Hper3(Ω)),\displaystyle\phi\in H^{1}(0,T;H^{-1}_{\mathrm{per}}(\Omega))\cap L^{\infty}(0,T;H^{2}_{\mathrm{per}}(\Omega))\cap L^{4}(0,T;H^{3}_{\mathrm{per}}(\Omega)),
μL2(0,T;Hper1(Ω)),\displaystyle\mu\in L^{2}(0,T;H^{1}_{\mathrm{per}}(\Omega)),
ωL(0,T;Lper2(Ω))L4(0,T;Hper1(Ω)),\displaystyle\omega\in L^{\infty}(0,T;L^{2}_{\mathrm{per}}(\Omega))\cap L^{4}(0,T;H^{1}_{\mathrm{per}}(\Omega)),
β(ϕ)L(0,T;Lper2(Ω))L4(0,T;Hper1(Ω)),\displaystyle\beta(\phi)\in L^{\infty}(0,T;L^{2}_{\mathrm{per}}(\Omega))\cap L^{4}(0,T;H^{1}_{\mathrm{per}}(\Omega)),
β(ϕ)β(ϕ)L2(0,T;Lper1(Ω)),β′′(ϕ)|ϕ|2L2(0,T;Lper1(Ω)),\displaystyle\beta(\phi)\beta^{\prime}(\phi)\in L^{2}(0,T;L^{1}_{\mathrm{per}}(\Omega)),\quad\beta^{\prime\prime}(\phi)|\nabla\phi|^{2}\in L^{2}(0,T;L^{1}_{\mathrm{per}}(\Omega)),
1<ϕ(x,t)<1a.e. in Ω×(0,T),\displaystyle-1<\phi(x,t)<1\quad\text{a.e.~{}in }\,\Omega\times(0,T),

and the initial condition is satisfied ϕ|t=0=ϕ0(x)\phi|_{t=0}=\phi_{0}(x) almost everywhere in Ω\Omega. The following weak formulations hold for almost all t(0,T)t\in(0,T):

tϕ=Δμin Hper1(Ω),\displaystyle\partial_{t}\phi=\Delta\mu\quad\text{in }\,H^{-1}_{\mathrm{per}}(\Omega), (2.3)
μ=ϵ4Δ2ϕ+β(ϕ)β(ϕ)+ϵ2β′′(ϕ)|ϕ|22ϵ2Δβ(ϕ)λϕβ(ϕ)+ϵ2(2λ+ϵpη)Δϕ\displaystyle\mu=\epsilon^{4}\Delta^{2}\phi+\beta(\phi)\beta^{\prime}(\phi)+\epsilon^{2}\beta^{\prime\prime}(\phi)|\nabla\phi|^{2}-2\epsilon^{2}\Delta\beta(\phi)-\lambda\phi\beta^{\prime}(\phi)+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta\phi
(λ+ϵpη)β(ϕ)+λ(λ+ϵpη)ϕ,in Hper1(Ω)+Lper1(Ω).\displaystyle\qquad-(\lambda+\epsilon^{p}\eta)\beta(\phi)+\lambda(\lambda+\epsilon^{p}\eta)\phi,\quad\text{in }\,H^{-1}_{\mathrm{per}}(\Omega)+L^{1}_{\mathrm{per}}(\Omega). (2.4)

Besides, the weak solution ϕ\phi satisfies the mass conservation such that

Ωϕ(x,t)dx=Ωϕ0(x)dx,t[0,T],\int_{\Omega}\phi(x,t)\,\mathrm{d}x=\int_{\Omega}\phi_{0}(x)\,\mathrm{d}x,\quad\forall\,t\in[0,T],

and for any t1,t2t_{1},t_{2} satisfying 0t1<t2T0\leq t_{1}<t_{2}\leq T, we have the energy equality:

E(ϕ(t2))+t1t2μ(s)L2(Ω)2ds=E(ϕ(t1)).E(\phi(t_{2}))+\int_{t_{1}}^{t_{2}}\|\nabla\mu(s)\|_{L^{2}(\Omega)}^{2}\,\mathrm{d}s=E(\phi(t_{1})).

Next, we present the property of strict separation from pure states ±1\pm 1 for global weak solutions. This property holds globally (in time) when the spatial dimension is less than or equal to two, while it can only hold locally in the three dimensional case. In analogy to [42, Proposition 2.8] and keeping in mind the continuous embedding H1(0,T;Hper1(Ω))L(0,T;Hper2(Ω))C([0,T];Cper(Ω¯))H^{1}(0,T;H^{-1}_{\mathrm{per}}(\Omega))\cap L^{\infty}(0,T;H^{2}_{\mathrm{per}}(\Omega))\hookrightarrow C([0,T];C_{\mathrm{per}}(\overline{\Omega})) for 1d31\leq d\leq 3 (cf. [57]), we can prove the following result:

Proposition 2.2 (Strict separation from pure states).

Let the hypotheses of Proposition 2.1 be satisfied.

(1)(1) Assume that d=1,2d=1,2 and T>0T>0. For any τ(0,T)\tau\in(0,T), there exists a constant δ(0,1)\delta\in(0,1) such that the global weak solution ϕ\phi obtained in Proposition 2.1 satisfies

ϕ(t)C(Ω¯)1δ,t[τ,T].\|\phi(t)\|_{C(\overline{\Omega})}\leq 1-\delta,\quad\forall\,t\in[\tau,\,T].

Moreover, if ϕ0C(Ω¯)1δ0\|\phi_{0}\|_{C(\overline{\Omega})}\leq 1-\delta_{0} for some given δ0(0,1)\delta_{0}\in(0,1), then there exists δ1(0,δ0]\delta_{1}\in(0,\delta_{0}] such that

ϕ(t)C(Ω¯)1δ1,t[0,T].\|\phi(t)\|_{C(\overline{\Omega})}\leq 1-\delta_{1},\quad\forall\,t\in[0,\,T].

(2)(2) Assume that d=3d=3 and ϕ0C(Ω¯)1δ0\|\phi_{0}\|_{C(\overline{\Omega})}\leq 1-\delta_{0} for some given δ0(0,1)\delta_{0}\in(0,1). Then there exists some T0(0,T]T_{0}\in(0,T] (depending on δ0\delta_{0}) such that

ϕ(t)C(Ω¯)112δ0,t[0,T0].\|\phi(t)\|_{C(\overline{\Omega})}\leq 1-\frac{1}{2}\delta_{0},\quad\forall\,t\in[0,\,T_{0}]. (2.5)
Remark 2.1.

Proposition 2.2 indicates that if the initial datum ϕ0\phi_{0} is strictly separated from ±1\pm 1, then in the one or two dimensional case, for an arbitrary but fixed final time T>0T>0, the solution ϕ\phi is strictly separated from ±1\pm 1 on the whole time interval [0,T][0,T] as well. On the other hand, in the three dimensional case, the strict separation property holds only in a “local” sense such that if ϕ0\phi_{0} is strictly separated from ±1\pm 1, then at least for some short period of time, the corresponding solution ϕ\phi can be strictly separated from ±1\pm 1. Whether a global in time separation property holds in the three dimensional case remains an open problem. The above separation property is crucial, since it enables us to avoid the singularity of the logarithmic type function β\beta (which is indeed smooth on a closed subset of (1,1)(-1,1)) and then obtain arbitrary higher order regularity of solutions to (1.1), provided that the initial datum is sufficiently smooth.

2.2 The finite difference spatial discretization

We first recall some standard notations for discrete functions and operators, see e.g., [58, 59]. For the spatial discretization, we apply the centered finite difference approximation and adopt the periodic setting similar to that in [28, 29, 55]. Without loss of generality, below we just state the setting in the computational domain Ω=(0,1)3\Omega=(0,1)^{3}. It can be easily extended to more general cells like Ω=(0,Lx)×(0,Ly)×(0,Lz)\Omega=(0,L_{x})\times(0,L_{y})\times(0,L_{z}) and to the lower dimensional cases d=1,2d=1,2.

Let NN\in\mathbb{N} be a given positive integer. We define the uniform spatial grid size as h=1/N>0h=1/N>0. The numerical value of a function ff at the cell centered mesh point ((i12)h,(j12)h,(k12)h)((i-\frac{1}{2})h,(j-\frac{1}{2})h,(k-\frac{1}{2})h) is denoted by fi,j,kf_{i,j,k} such that fi,j,k=f((i12)h,(j12)h,(k12)h)f_{i,j,k}=f((i-\frac{1}{2})h,(j-\frac{1}{2})h,(k-\frac{1}{2})h). Then we introduce the space of three dimensional discrete N3N^{3}-periodic functions:

𝒞per:={f:3|fi,j,k=fi+γ1N,j+γ2N,k+γ3N,i,j,k,γ1,γ2,γ3}.\mathcal{C}_{\mathrm{per}}:=\left\{f:\mathbb{Z}^{3}\rightarrow\mathbb{R}\,|\,f_{i,j,k}=f_{i+\gamma_{1}N,j+\gamma_{2}N,k+\gamma_{3}N},\ \ \forall\,i,j,k,\gamma_{1},\gamma_{2},\gamma_{3}\in\mathbb{Z}\right\}.

The discrete inner product on 𝒞per\mathcal{C}_{\mathrm{per}} is defined as f,gΩ:=h3i,j,k=1Nfi,j,kgi,j,k\left<f,g\right>_{\Omega}:=h^{3}\sum_{i,j,k=1}^{N}f_{i,j,k}g_{i,j,k} for any f,g𝒞perf,g\in\mathcal{C}_{\mathrm{per}}. For any f𝒞perf\in\mathcal{C}_{\mathrm{per}}, its mean value is given by f¯=|Ω|1f,1Ω\overline{f}=|\Omega|^{-1}\left<f,1\right>_{\Omega} (here we simply have |Ω|=1|\Omega|=1). Then we introduce the subspace of 𝒞per\mathcal{C}_{\mathrm{per}} with zero mean such that

𝒞̊per:={f𝒞per|f¯=0}.\mathring{\mathcal{C}}_{\mathrm{per}}:=\left\{f\in\mathcal{C}_{\mathrm{per}}\,|\,\overline{f}=0\right\}.

Next, we use the notation per=perx×pery×perz\mathcal{E}_{\mathrm{per}}=\mathcal{E}_{\mathrm{per}}^{x}\times\mathcal{E}_{\mathrm{per}}^{y}\times\mathcal{E}_{\mathrm{per}}^{z} to denote the set of face-centered vector grid functions f=(fx,fy,fz)T\vec{f}=(f^{x},f^{y},f^{z})^{T} that are Ω\Omega-periodic, i.e., NN-periodic in each direction. Here, we set

perx:={f:3|fi+12,j,k=fi+12+γ1N,j+γ2N,k+γ3N,i,j,k,γ1,γ2,γ3},\mathcal{E}_{\mathrm{per}}^{x}:=\left\{f:\mathbb{Z}^{3}\rightarrow\mathbb{R}\,\Big{|}\,f_{i+\frac{1}{2},j,k}=f_{i+\frac{1}{2}+\gamma_{1}N,j+\gamma_{2}N,k+\gamma_{3}N},\ \ \forall\,i,j,k,\gamma_{1},\gamma_{2},\gamma_{3}\in\mathbb{Z}\right\},

where fi+12,j,k=f(ih,(j12)h,(k12)h)f_{i+\frac{1}{2},j,k}=f(ih,(j-\frac{1}{2})h,(k-\frac{1}{2})h). The other two spaces pery\mathcal{E}_{\mathrm{per}}^{y}, perz\mathcal{E}_{\mathrm{per}}^{z} are defined in an analogous manner.

The discrete average and difference operators acting on 𝒞per\mathcal{C}_{\mathrm{per}} yield face-centered functions, evaluated at the east-west faces, (i+12,j,k)(i+\frac{1}{2},j,k); north-south faces, (i,j+12,k)(i,j+\frac{1}{2},k); and up-down faces, (i,j,k+12)(i,j,k+\frac{1}{2}), respectively. We define Ax,Dx:𝒞perperxA_{x},D_{x}:\mathcal{C}_{\mathrm{per}}\to\mathcal{E}_{\mathrm{per}}^{x}, Ay,Dy:𝒞perperyA_{y},D_{y}:\mathcal{C}_{\mathrm{per}}\to\mathcal{E}_{\mathrm{per}}^{y} and Az,Dz:𝒞perperzA_{z},D_{z}:\mathcal{C}_{\mathrm{per}}\to\mathcal{E}_{\mathrm{per}}^{z} such that

Axfi+12,j,k:=12(fi+1,j,k+fi,j,k),Dxfi+12,j,k:=1h(fi+1,j,kfi,j,k),\displaystyle A_{x}f_{i+\frac{1}{2},j,k}:=\frac{1}{2}\left(f_{i+1,j,k}+f_{i,j,k}\right),\qquad D_{x}f_{i+\frac{1}{2},j,k}:=\frac{1}{h}\left(f_{i+1,j,k}-f_{i,j,k}\right),
Ayfi,j+12,k:=12(fi,j+1,k+fi,j,k),Dyfi,j+12,k:=1h(fi,j+1,kfi,j,k),\displaystyle A_{y}f_{i,j+\frac{1}{2},k}:=\frac{1}{2}\left(f_{i,j+1,k}+f_{i,j,k}\right),\qquad D_{y}f_{i,j+\frac{1}{2},k}:=\frac{1}{h}\left(f_{i,j+1,k}-f_{i,j,k}\right),
Azfi,j,k+12:=12(fi,j,k+1+fi,j,k),Dzfi,j,k+12:=1h(fi,j,k+1fi,j,k).\displaystyle A_{z}f_{i,j,k+\frac{1}{2}}:=\frac{1}{2}\left(f_{i,j,k+1}+f_{i,j,k}\right),\qquad D_{z}f_{i,j,k+\frac{1}{2}}:=\frac{1}{h}\left(f_{i,j,k+1}-f_{i,j,k}\right).

On the other hand, the corresponding operators at the staggered mesh points, that is, ax,dx:perx𝒞pera_{x},d_{x}:\mathcal{E}_{\mathrm{per}}^{x}\to\mathcal{C}_{\mathrm{per}}, ay,dy:pery𝒞pera_{y},d_{y}:\mathcal{E}_{\mathrm{per}}^{y}\to\mathcal{C}_{\mathrm{per}}, az,dz:perz𝒞pera_{z},d_{z}:\mathcal{E}_{\mathrm{per}}^{z}\to\mathcal{C}_{\mathrm{per}}, are given by

axfi,j,k:=12(fi+12,j,k+fi12,j,k),dxfi,j,k:=1h(fi+12,j,kfi12,j,k),\displaystyle a_{x}f_{i,j,k}:=\frac{1}{2}\left(f_{i+\frac{1}{2},j,k}+f_{i-\frac{1}{2},j,k}\right),\qquad d_{x}f_{i,j,k}:=\frac{1}{h}\left(f_{i+\frac{1}{2},j,k}-f_{i-\frac{1}{2},j,k}\right),
ayfi,j,k:=12(fi,j+12,k+fi,j12,k),dyfi,j,k:=1h(fi,j+12,kfi,j12,k),\displaystyle a_{y}f_{i,j,k}:=\frac{1}{2}\left(f_{i,j+\frac{1}{2},k}+f_{i,j-\frac{1}{2},k}\right),\qquad d_{y}f_{i,j,k}:=\frac{1}{h}\left(f_{i,j+\frac{1}{2},k}-f_{i,j-\frac{1}{2},k}\right),
azfi,j,k:=12(fi,j,k+12+fi,j,k12),dzfi,j,k:=1h(fi,j,k+12fi,j,k12).\displaystyle a_{z}f_{i,j,k}:=\frac{1}{2}\left(f_{i,j,k+\frac{1}{2}}+f_{i,j,k-\frac{1}{2}}\right),\qquad d_{z}f_{i,j,k}:=\frac{1}{h}\left(f_{i,j,k+\frac{1}{2}}-f_{i,j,k-\frac{1}{2}}\right).

Let us now introduce the discrete gradient and divergence operators h:𝒞perper\nabla_{h}:\mathcal{C}_{\mathrm{per}}\to\mathcal{E}_{\mathrm{per}}, h:per𝒞per\nabla_{h}\cdot:\mathcal{E}_{\mathrm{per}}\to\mathcal{C}_{\mathrm{per}}. For any f𝒞perf\in\mathcal{C}_{\mathrm{per}}, we set hfi,j,k=(Dxfi+12,j,k,Dyfi,j+12,k,Dzfi,j,k+12)T\nabla_{h}f_{i,j,k}=(D_{x}f_{i+\frac{1}{2},j,k},D_{y}f_{i,j+\frac{1}{2},k},D_{z}f_{i,j,k+\frac{1}{2}})^{T}, while for any f=(fx,fy,fz)Tper\vec{f}=(f^{x},f^{y},f^{z})^{T}\in\mathcal{E}_{\mathrm{per}}, we define hfi,j,k=dxfi,j,kx+dyfi,j,ky+dzfi,j,kz\nabla_{h}\cdot\vec{f}_{i,j,k}=d_{x}f^{x}_{i,j,k}+d_{y}f^{y}_{i,j,k}+d_{z}f^{z}_{i,j,k}. Suppose that gg is an arbitrary scalar function defined at all the face center points (east-west, north-south, and up-down) and f=(fx,fy,fz)Tper\vec{f}=(f^{x},f^{y},f^{z})^{T}\in\mathcal{E}_{\mathrm{per}}, then we have

h(gf)i,j,k=dx(gewfx)i,j,k+dy(gnsfy)i,j,k+dz(gudfz)i,j,k.\nabla_{h}\cdot(g\vec{f})_{i,j,k}=d_{x}(g_{\text{ew}}f^{x})_{i,j,k}+d_{y}(g_{\text{ns}}f^{y})_{i,j,k}+d_{z}(g_{\text{ud}}f^{z})_{i,j,k}.

If f=hϕ\vec{f}=\nabla_{h}\phi for certain scalar grid function ϕ𝒞per\phi\in\mathcal{C}_{\mathrm{per}}, it follows that

h(ghϕ)i,j,k=dx(gewDxϕ)i,j,k+dy(gnsDyϕ)i,j,k+dz(gudDzϕ)i,j,k.\nabla_{h}\cdot\left(g\nabla_{h}\phi\right)_{i,j,k}=d_{x}(g_{\text{ew}}D_{x}\phi)_{i,j,k}+d_{y}(g_{\text{ns}}D_{y}\phi)_{i,j,k}+d_{z}(g_{\text{ud}}D_{z}\phi)_{i,j,k}.

As a consequence, the standard three dimensional discrete Laplacian Δh:𝒞per𝒞per\Delta_{h}:\mathcal{C}_{\mathrm{per}}\rightarrow\mathcal{C}_{\mathrm{per}} can be defined as

Δhϕ:=h(1hϕ)i,j,k=dx(Dxϕ)i,j,k+dy(Dyϕ)i,j,k+dz(Dzϕ)i,j,k,ϕ𝒞per.\Delta_{h}\phi:=\nabla_{h}\cdot(1\nabla_{h}\phi)_{i,j,k}=d_{x}(D_{x}\phi)_{i,j,k}+d_{y}(D_{y}\phi)_{i,j,k}+d_{z}(D_{z}\phi)_{i,j,k},\quad\forall\,\phi\in\mathcal{C}_{\mathrm{per}}.

For 1p<1\leq p<\infty, the discrete lpl^{p} norms are given by fLhp:=(|f|p,1Ω)1p\|f\|_{L_{h}^{p}}:=\left(\left<|f|^{p},1\right>_{\Omega}\right)^{\frac{1}{p}}. When p=2p=2, we simply denote fLh2\|f\|_{L_{h}^{2}} by f2\|f\|_{2}. In addition, the discrete maximum norm in LhL^{\infty}_{h} is defined as f:=max1i,j,kN|fi,j,k|\|f\|_{\infty}:=\max_{1\leq i,j,k\leq N}|f_{i,j,k}|. For any two vector grid functions f=(fx,fy,fz)T,g=(gx,gy,gz)Tper\vec{f}=(f^{x},f^{y},f^{z})^{T},\vec{g}=(g^{x},g^{y},g^{z})^{T}\in\mathcal{E}_{\mathrm{per}}, we define the discrete inner product as follows

[f,g]Ω:=[fx,gx]x+[fy,gy]y+[fz,gz]z,\big{[}\vec{f},\vec{g}\big{]}_{\Omega}:=\left[f^{x},g^{x}\right]_{x}+\left[f^{y},g^{y}\right]_{y}+\left[f^{z},g^{z}\right]_{z},

where

[fx,gx]x:=ax(fxgx),1Ω,[fy,gy]y:=ay(fygy),1Ω,[fz,gz]z:=az(fzgz),1Ω.\left[f^{x},g^{x}\right]_{x}:=\left<a_{x}(f^{x}g^{x}),1\right>_{\Omega},\quad\left[f^{y},g^{y}\right]_{y}:=\left<a_{y}(f^{y}g^{y}),1\right>_{\Omega},\quad\left[f^{z},g^{z}\right]_{z}:=\left<a_{z}(f^{z}g^{z}),1\right>_{\Omega}.

With the above notations, we can further define

hfLhp:=([|Dxf|p,1]x+[|Dyf|p,1]y+[|Dzf|p,1]z)1p,1p<,\displaystyle\|\nabla_{h}f\|_{L_{h}^{p}}:=\left(\big{[}|D_{x}f|^{p},1\big{]}_{x}+\big{[}|D_{y}f|^{p},1\big{]}_{y}+\big{[}|D_{z}f|^{p},1\big{]}_{z}\right)^{\frac{1}{p}},\quad 1\leq p<\infty, (2.6)

and the discrete Sobolev-type norms through

fWh1,pp:=fLhpp+hfLhpp,fWh2,pp:=fWh1,pp+ΔhfLhpp,f𝒞per,\|f\|_{W_{h}^{1,p}}^{p}:=\|f\|_{L_{h}^{p}}^{p}+\|\nabla_{h}f\|_{L_{h}^{p}}^{p},\quad\|f\|_{W_{h}^{2,p}}^{p}:=\|f\|_{W_{h}^{1,p}}^{p}+\|\Delta_{h}f\|_{L_{h}^{p}}^{p},\quad\forall\,f\in\mathcal{C}_{\mathrm{per}},

and so on. When p=2p=2, we simply denote the norm Whm,2\|\cdot\|_{W_{h}^{m,2}} by Hhm\|\cdot\|_{H_{h}^{m}}, m+m\in\mathbb{Z}^{+}.

A discrete version of the space H̊per1(Ω)\mathring{H}^{-1}_{\mathrm{per}}(\Omega) is also necessary in the subsequent analysis. For any f𝒞̊perf\in\mathring{\mathcal{C}}_{\mathrm{per}}, let ψ[f]:=(Δh)1f𝒞̊per\psi[f]:=(-\Delta_{h})^{-1}f\in\mathring{\mathcal{C}}_{\mathrm{per}} be the unique periodic solution of the discrete elliptic equation Δhψ=f-\Delta_{h}\psi=f (cf. [59, Lemma 3.2]). Then the inner product ,1,h\left<\cdot,\cdot\right>_{-1,h} and the associated 1,h\|\cdot\|_{-1,h} norm can be defined by

f,g1,h\displaystyle\left<f,g\right>_{-1,h} :=[hψ[f],hψ[g]]Ω=f,ψ[g]Ω=ψ[f],gΩ,f,g𝒞̊per,\displaystyle:=\big{[}\nabla_{h}\psi[f],\nabla_{h}\psi[g]\big{]}_{\Omega}=\left<f,\psi[g]\right>_{\Omega}=\left<\psi[f],g\right>_{\Omega},\quad\forall\,f,g\in\mathring{\mathcal{C}}_{\mathrm{per}},

and f1,h2:=f,f1,h\|f\|^{2}_{-1,h}:=\left<f,f\right>_{-1,h} for any f𝒞̊perf\in\mathring{\mathcal{C}}_{\mathrm{per}}. The above definition is a direct consequence of the following lemma on the discrete summation-by-parts formulae, cf. [59, Lemma 3.3], see also [58, 60, 61].

Lemma 2.1.

Let 𝒟\mathcal{D} be an arbitrary periodic scalar function defined on all of the face center points. For any ψ,ϕ𝒞per\psi,\phi\in\mathcal{C}_{\mathrm{per}} and any fper\vec{f}\in\mathcal{E}_{\mathrm{per}}, the following summation-by-parts formulae are valid:

<ψ,hf>Ω=[hψ,f]Ω,<ψ,h(𝒟hϕ)>Ω=[hψ,𝒟hϕ]Ω.\big{<}\psi,\nabla_{h}\cdot\vec{f}\,\big{>}_{\Omega}=-\big{[}\nabla_{h}\psi,\vec{f}\,\big{]}_{\Omega},\qquad\big{<}\psi,\nabla_{h}\cdot(\mathcal{D}\nabla_{h}\phi)\big{>}_{\Omega}=-\big{[}\nabla_{h}\psi,\mathcal{D}\nabla_{h}\phi\big{]}_{\Omega}.

Before ending this subsection, we also recall the following result (see [29, Lemma 3.1]) that will be useful in the subsequent analysis:

Lemma 2.2.

Suppose that ϕ1,ϕ2𝒞per\phi_{1},\phi_{2}\in\mathcal{C}_{\mathrm{per}}, satisfying ϕ1ϕ2,1Ω=0\left<\phi_{1}-\phi_{2},1\right>_{\Omega}=0 and ϕ1,ϕ2<1\|\phi_{1}\|_{\infty},\,\|\phi_{2}\|_{\infty}<1, the following estimate holds:

(Δh)1(ϕ1ϕ2)C0,\|(-\Delta_{h})^{-1}(\phi_{1}-\phi_{2})\|_{\infty}\leq C_{0}, (2.7)

where the constant C0>0C_{0}>0 only depends on Ω\Omega.

2.3 A convex-concave energy decomposition

Before proceeding, let us mention that although the sign of the parameters λ,η\lambda,\eta does not play an essential role in the theoretic analysis of the continuous problem (1.1), it will leads to some differences in the corresponding numerical schemes (see Remark 2.3 for detailed discussions). In the remaining part of this paper, we will focus on the physical relevant case (i.e., the FCH energy with a logarithmic, possibly double-well potential FF) and assume that

λ,η,ϵ>0.\lambda,\ \eta,\ \epsilon>0.

In order to design a numerical scheme with unconditional energy stability, we shall perform a convex-concave decomposition for the FCH energy given by (1.2). In view of (2.1) and (2.2), it can be rewritten in the following way:

E(ϕ)\displaystyle E(\phi) =ϵ42ΔϕL2(Ω)2+12β(ϕ)L2(Ω)2+(λ22+λϵpη2)ϕL2(Ω)2ϵpηΩB(ϕ)dx\displaystyle=\frac{\epsilon^{4}}{2}\|\Delta\phi\|_{L^{2}(\Omega)}^{2}+\frac{1}{2}\|\beta(\phi)\|_{L^{2}(\Omega)}^{2}+\left(\frac{\lambda^{2}}{2}+\frac{\lambda\epsilon^{p}\eta}{2}\right)\|\phi\|_{L^{2}(\Omega)}^{2}-\epsilon^{p}\eta\int_{\Omega}B(\phi)\,\mathrm{d}x
(ϵ2+pη2+λϵ2)ϕL2(Ω)2λΩϕβ(ϕ)dxϵ2Ωβ(ϕ)Δϕdx.\displaystyle\quad-\left(\frac{\epsilon^{2+p}\eta}{2}+\lambda\epsilon^{2}\right)\|\nabla\phi\|_{L^{2}(\Omega)}^{2}-\lambda\int_{\Omega}\phi\beta(\phi)\,\mathrm{d}x-\epsilon^{2}\int_{\Omega}\beta(\phi)\Delta\phi\,\mathrm{d}x. (2.8)

The term Ωβ(ϕ)Δϕdx-\int_{\Omega}\beta(\phi)\Delta\phi\,\mathrm{d}x in (2.8) turns out to be tricky since it is neither convex nor concave. To overcome this difficulty, using the periodic boundary conditions, we perform integration by parts and obtain

E(ϕ)\displaystyle E(\phi) =ϵ42ΔϕL2(Ω)2+12β(ϕ)L2(Ω)2+(λ22+λϵpη2)ϕL2(Ω)2ϵpηΩB(ϕ)dx\displaystyle=\frac{\epsilon^{4}}{2}\|\Delta\phi\|_{L^{2}(\Omega)}^{2}+\frac{1}{2}\|\beta(\phi)\|_{L^{2}(\Omega)}^{2}+\left(\frac{\lambda^{2}}{2}+\frac{\lambda\epsilon^{p}\eta}{2}\right)\|\phi\|_{L^{2}(\Omega)}^{2}-\epsilon^{p}\eta\int_{\Omega}B(\phi)\,\mathrm{d}x
(ϵ2+pη2+λϵ2)ϕL2(Ω)2λΩϕβ(ϕ)dx+ϵ2Ωβ(ϕ)|ϕ|2dx.\displaystyle\quad-\left(\frac{\epsilon^{2+p}\eta}{2}+\lambda\epsilon^{2}\right)\|\nabla\phi\|_{L^{2}(\Omega)}^{2}-\lambda\int_{\Omega}\phi\beta(\phi)\,\mathrm{d}x+\epsilon^{2}\int_{\Omega}\beta^{\prime}(\phi)|\nabla\phi|^{2}\,\mathrm{d}x. (2.9)

The above procedure is meaningful, if the function ϕ\phi is sufficiently regular, for example, ϕHper2(Ω)\phi\in H^{2}_{\mathrm{per}}(\Omega) and satisfies the strict separation property (keeping in mind the singular nature of β\beta and its derivatives).

Thanks to (2.9), we can derive the following result:

Proposition 2.3.

The reformulated FCH energy (2.9) possesses a convex-concave decomposition such that

E(ϕ)=Ec(ϕ)Ee(ϕ),E(\phi)=E^{c}(\phi)-E^{e}(\phi),

where

Ec(ϕ)=ϵ42ΔϕL2(Ω)2+12β(ϕ)L2(Ω)2+(λ22+λϵpη2)ϕL2(Ω)2+ϵ2Ωβ(ϕ)|ϕ|2dx,Ee(ϕ)=(ϵ2+pη2+λϵ2)ϕL2(Ω)2+λΩϕβ(ϕ)dx+ϵpηΩB(ϕ)dx.\displaystyle\begin{aligned} &E^{c}(\phi)=\frac{\epsilon^{4}}{2}\|\Delta\phi\|_{L^{2}(\Omega)}^{2}+\frac{1}{2}\|\beta(\phi)\|_{L^{2}(\Omega)}^{2}+\left(\frac{\lambda^{2}}{2}+\frac{\lambda\epsilon^{p}\eta}{2}\right)\|\phi\|_{L^{2}(\Omega)}^{2}+\epsilon^{2}\int_{\Omega}\beta^{\prime}(\phi)|\nabla\phi|^{2}\,\mathrm{d}x,\\ &E^{e}(\phi)=\left(\frac{\epsilon^{2+p}\eta}{2}+\lambda\epsilon^{2}\right)\|\nabla\phi\|_{L^{2}(\Omega)}^{2}+\lambda\int_{\Omega}\phi\beta(\phi)\,\mathrm{d}x+\epsilon^{p}\eta\int_{\Omega}B(\phi)\,\mathrm{d}x.\end{aligned}

For any given δ(0,1)\delta\in(0,1), both Ec(ϕ)E^{c}(\phi) and Ee(ϕ)E^{e}(\phi) are strictly convex when restricted to the set of functions ϕHper2(Ω)\phi\in H^{2}_{\mathrm{per}}(\Omega) satisfying |ϕ(x)|1δ|\phi(x)|\leq 1-\delta in Ω\Omega.

Proof.

We adapt the arguments in [42, Lemma 5.3]. For any ϕHper2(Ω)\phi\in H^{2}_{\mathrm{per}}(\Omega) satisfying |ϕ(x)|1δ|\phi(x)|\leq 1-\delta for some δ(0,1)\delta\in(0,1), by a direct calculation we can deduce that

(Ec)(ϕ),v\displaystyle\left<(E^{c})^{\prime}(\phi),v\right> =ϵ4ΩΔϕΔvdx+Ωβ(ϕ)β(ϕ)vdx+(λ2+λϵpη)Ωϕvdx\displaystyle=\epsilon^{4}\int_{\Omega}\Delta\phi\Delta v\,\mathrm{d}x+\int_{\Omega}\beta(\phi)\beta^{\prime}(\phi)v\,\mathrm{d}x+\left(\lambda^{2}+\lambda\epsilon^{p}\eta\right)\int_{\Omega}\phi v\,\mathrm{d}x
+ϵ2Ω(2β(ϕ)ϕv+β′′(ϕ)|ϕ|2v)dx,\displaystyle\quad+\epsilon^{2}\int_{\Omega}\big{(}2\beta^{\prime}(\phi)\nabla\phi\cdot\nabla v+\beta^{\prime\prime}(\phi)|\nabla\phi|^{2}v\big{)}\,\mathrm{d}x,
(Ee)(ϕ),v\displaystyle\left<(E^{e})^{\prime}(\phi),v\right> =(ϵ2+pη+2λϵ2)Ωϕvdx+λΩ(β(ϕ)+ϕβ(ϕ))vdx+ϵpηΩβ(ϕ)vdx,\displaystyle=\left(\epsilon^{2+p}\eta+2\lambda\epsilon^{2}\right)\int_{\Omega}\nabla\phi\cdot\nabla v\,\mathrm{d}x+\lambda\int_{\Omega}(\beta(\phi)+\phi\beta^{\prime}(\phi))v\,\mathrm{d}x+\epsilon^{p}\eta\int_{\Omega}\beta(\phi)v\,\mathrm{d}x,

for any vHper2(Ω)v\in H^{2}_{\mathrm{per}}(\Omega). A further calculation yields

(Ec)′′(ϕ)v,w\displaystyle\left<(E^{c})^{\prime\prime}(\phi)v,w\right> =ϵ4ΩΔvΔwdx+Ω[(β(ϕ))2+β(ϕ)β′′(ϕ)]vwdx+(λ2+λϵpη)Ωvwdx\displaystyle=\epsilon^{4}\int_{\Omega}\Delta v\Delta w\,\mathrm{d}x+\int_{\Omega}\left[(\beta^{\prime}(\phi))^{2}+\beta(\phi)\beta^{\prime\prime}(\phi)\right]vw\,\mathrm{d}x+\left(\lambda^{2}+\lambda\epsilon^{p}\eta\right)\int_{\Omega}vw\,\mathrm{d}x
+ϵ2Ω(2β(ϕ)vw+2β′′(ϕ)vϕw)dx\displaystyle\quad+\epsilon^{2}\int_{\Omega}\big{(}2\beta^{\prime}(\phi)\nabla v\cdot\nabla w+2\beta^{\prime\prime}(\phi)v\nabla\phi\cdot\nabla w\big{)}\,\mathrm{d}x
+ϵ2Ω(2β′′(ϕ)wϕv+β′′′(ϕ)|ϕ|2vw)dx,\displaystyle\quad+\epsilon^{2}\int_{\Omega}\big{(}2\beta^{\prime\prime}(\phi)w\nabla\phi\cdot\nabla v+\beta^{\prime\prime\prime}(\phi)|\nabla\phi|^{2}vw\big{)}\,\mathrm{d}x,
(Ee)′′(ϕ)v,w\displaystyle\left<(E^{e})^{\prime\prime}(\phi)v,w\right> =(ϵ2+pη+2λϵ2)Ωvwdx+Ω[(2λ+ϵpη)β(ϕ)+λϕβ′′(ϕ)]vwdx,\displaystyle=(\epsilon^{2+p}\eta+2\lambda\epsilon^{2})\int_{\Omega}\nabla v\cdot\nabla w\,\mathrm{d}x+\int_{\Omega}\big{[}\left(2\lambda+\epsilon^{p}\eta\right)\beta^{\prime}(\phi)+\lambda\phi\beta^{\prime\prime}(\phi)\big{]}vw\,\mathrm{d}x,

for any v,wHper2(Ω)v,w\in H^{2}_{\mathrm{per}}(\Omega). Then taking w=vw=v, we find that (cf. [50, (6.5)])

(Ec)′′(ϕ)v,v\displaystyle\left<(E^{c})^{\prime\prime}(\phi)v,v\right> ϵ4Ω|Δv|2dx+Ω[(β(ϕ))2+β(ϕ)β′′(ϕ)]v2dx+(λ2+λϵpη)Ωv2dx\displaystyle\geq\epsilon^{4}\int_{\Omega}|\Delta v|^{2}\,\mathrm{d}x+\int_{\Omega}\left[(\beta^{\prime}(\phi))^{2}+\beta(\phi)\beta^{\prime\prime}(\phi)\right]v^{2}\,\mathrm{d}x+\left(\lambda^{2}+\lambda\epsilon^{p}\eta\right)\int_{\Omega}v^{2}\,\mathrm{d}x
+ϵ2Ω[2β(ϕ)4(β′′(ϕ))2β′′′(ϕ)]|v|2dx,\displaystyle\quad+\epsilon^{2}\int_{\Omega}\left[2\beta^{\prime}(\phi)-\frac{4(\beta^{\prime\prime}(\phi))^{2}}{\beta^{\prime\prime\prime}(\phi)}\right]|\nabla v|^{2}\,\mathrm{d}x,
(Ee)′′(ϕ)v,v\displaystyle\left<(E^{e})^{\prime\prime}(\phi)v,v\right> =(ϵ2+pη+2λϵ2)Ω|v|2dx+Ω[(2λ+ϵpη)β(ϕ)+λϕβ′′(ϕ)]v2dx.\displaystyle=(\epsilon^{2+p}\eta+2\lambda\epsilon^{2})\int_{\Omega}|\nabla v|^{2}\,\mathrm{d}x+\int_{\Omega}\big{[}\left(2\lambda+\epsilon^{p}\eta\right)\beta^{\prime}(\phi)+\lambda\phi\beta^{\prime\prime}(\phi)\big{]}v^{2}\,\mathrm{d}x.

Recalling (2.2), we observe that for r(1,1)r\in(-1,1), it holds

β(r)=21r22,\displaystyle\beta^{\prime}(r)=\frac{2}{1-r^{2}}\geq 2, (2.10)
β(r)β′′(r)=ln1+r1r4r(1r2)20,\displaystyle\beta(r)\beta^{\prime\prime}(r)=\ln\frac{1+r}{1-r}\frac{4r}{(1-r^{2})^{2}}\geq 0, (2.11)
β(r)2(β′′(r))2β′′′(r)=21+3r2>12,\displaystyle\beta^{\prime}(r)-\frac{2(\beta^{\prime\prime}(r))^{2}}{\beta^{\prime\prime\prime}(r)}=\frac{2}{1+3r^{2}}>\frac{1}{2}, (2.12)
rβ′′(r)=4r2(1r2)20.\displaystyle r\beta^{\prime\prime}(r)=\frac{4r^{2}}{(1-r^{2})^{2}}\geq 0. (2.13)

As a consequence, we can conclude

(Ec)′′(ϕ)v,vCvH2(Ω)2and(Ee)′′(ϕ)v,vCvH1(Ω)2,vHper2(Ω),\left<(E^{c})^{\prime\prime}(\phi)v,v\right>\geq C\|v\|_{H^{2}(\Omega)}^{2}\quad\text{and}\quad\left<(E^{e})^{\prime\prime}(\phi)v,v\right>\geq C\|v\|_{H^{1}(\Omega)}^{2},\quad\forall\,v\in H^{2}_{\mathrm{per}}(\Omega),

where C>0C>0 is a constant that may depend on the positive parameters λ\lambda, η\eta, ϵ\epsilon and δ\delta, but are independent of ϕ\phi and vv. The proof is complete. ∎

Corresponding to (2.9), for any h>0h>0, we introduce the discrete energy

Eh(ϕ)\displaystyle E_{h}(\phi) =ϵ42Δhϕ22+12β(ϕ)22+(λ22+λϵpη2)ϕ22(ϵ2+pη2+λϵ2)hϕ22\displaystyle=\frac{\epsilon^{4}}{2}\|\Delta_{h}\phi\|_{2}^{2}+\frac{1}{2}\|\beta(\phi)\|_{2}^{2}+\left(\frac{\lambda^{2}}{2}+\frac{\lambda\epsilon^{p}\eta}{2}\right)\|\phi\|_{2}^{2}-\left(\frac{\epsilon^{2+p}\eta}{2}+\lambda\epsilon^{2}\right)\|\nabla_{h}\phi\|_{2}^{2}
λϕ,β(ϕ)Ω+ϵ2β(ϕ),ζ=x,y,zaζ(|Dζϕ|2)ΩϵpηB(ϕ),1Ω.\displaystyle\quad-\lambda\left<\phi,\beta(\phi)\right>_{\Omega}+\epsilon^{2}\left<\beta^{\prime}(\phi),\sum_{\zeta=x,y,z}a_{\zeta}\left(\left|D_{\zeta}\phi\right|^{2}\right)\right>_{\Omega}-\epsilon^{p}\eta\left<B(\phi),1\right>_{\Omega}. (2.14)

In analogy to Proposition 2.3, we can deduce that

Corollary 2.1.

The discrete energy EhE_{h} possesses a convex-concave decomposition:

Eh(ϕ)=Ehc(ϕ)Ehe(ϕ),E_{h}(\phi)=E_{h}^{c}(\phi)-E_{h}^{e}(\phi), (2.15)

where

Ehc(ϕ)=ϵ42Δhϕ22+12β(ϕ)22+(λ22+λϵpη2)ϕ22+ϵ2β(ϕ),ζ=x,y,zaζ(|Dζϕ|2)Ω,\displaystyle E_{h}^{c}(\phi)=\frac{\epsilon^{4}}{2}\|\Delta_{h}\phi\|_{2}^{2}+\frac{1}{2}\|\beta(\phi)\|_{2}^{2}+\left(\frac{\lambda^{2}}{2}+\frac{\lambda\epsilon^{p}\eta}{2}\right)\|\phi\|_{2}^{2}+\epsilon^{2}\left<\beta^{\prime}(\phi),\sum_{\zeta=x,y,z}a_{\zeta}\left(\left|D_{\zeta}\phi\right|^{2}\right)\right>_{\Omega}, (2.16)
Ehe(ϕ)=(ϵ2+pη2+λϵ2)hϕ22+λϕ,β(ϕ)Ω+ϵpηB(ϕ),1Ω.\displaystyle E_{h}^{e}(\phi)=\left(\frac{\epsilon^{2+p}\eta}{2}+\lambda\epsilon^{2}\right)\|\nabla_{h}\phi\|_{2}^{2}+\lambda\left<\phi,\beta(\phi)\right>_{\Omega}+\epsilon^{p}\eta\left<B(\phi),1\right>_{\Omega}. (2.17)

Both EhcE_{h}^{c} and EheE_{h}^{e} are strictly convex when restricted to the set of functions ϕ𝒞per\phi\in\mathcal{C}_{\mathrm{per}} satisfying ϕ<1\|\phi\|_{\infty}<1.

Remark 2.2.

Since ϕ𝒞per\phi\in\mathcal{C}_{\mathrm{per}} only takes its values at finite points in Ω\Omega, then the condition ϕ<1\|\phi\|_{\infty}<1 indeed implies that there exists some δ(0,1)\delta\in(0,1) such that ϕ1δ\|\phi\|_{\infty}\leq 1-\delta.

2.4 The first order numerical scheme

Using the convex splitting method [51, 52] and the convex-concave energy decomposition in Corollary 2.1, we introduce a first order semi-implicit numerical scheme for the functionalized Cahn-Hilliard equation (1.1) (cf. also (2.4) for an alternative expression of μ\mu). To this end, for a uniform time step Δt>0\Delta t>0, given ϕn𝒞per\phi^{n}\in\mathcal{C}_{\mathrm{per}} with ϕn<1\|\phi^{n}\|_{\infty}<1 (nn\in\mathbb{N}), we look for the discrete solution ϕn+1\phi^{n+1} such that

ϕn+1ϕnΔt=Δhμn+1,\displaystyle\frac{\phi^{n+1}-\phi^{n}}{\Delta t}=\Delta_{h}\mu^{n+1}, (2.18)
μn+1=ϵ4Δh2ϕn+1+β(ϕn+1)β(ϕn+1)+ϵ2ζ=x,y,zβ′′(ϕn+1)aζ(|Dζϕn+1|2)\displaystyle\mu^{n+1}=\epsilon^{4}\Delta_{h}^{2}\phi^{n+1}+\beta(\phi^{n+1})\beta^{\prime}(\phi^{n+1})+\epsilon^{2}\sum_{\zeta=x,y,z}\beta^{\prime\prime}(\phi^{n+1})a_{\zeta}\left(|D_{\zeta}\phi^{n+1}|^{2}\right)
2ϵ2ζ=x,y,zdζ(Aζ(β(ϕn+1))Dζϕn+1)+λ(λ+ϵpη)ϕn+1\displaystyle\qquad\quad-2\epsilon^{2}\sum_{\zeta=x,y,z}d_{\zeta}\left(A_{\zeta}\big{(}\beta^{\prime}(\phi^{n+1})\big{)}D_{\zeta}\phi^{n+1}\right)+\lambda(\lambda+\epsilon^{p}\eta)\phi^{n+1}
+ϵ2(2λ+ϵpη)Δhϕnλϕnβ(ϕn)(λ+ϵpη)β(ϕn).\displaystyle\qquad\quad+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta_{h}\phi^{n}-\lambda\phi^{n}\beta^{\prime}(\phi^{n})-(\lambda+\epsilon^{p}\eta)\beta(\phi^{n}). (2.19)

It is straightforward to check that for every solution to the numerical scheme (2.18)–(2.19), if it exists, then it must satisfy the property of mass conservation (cf. Proposition 2.1 for the continuous problem). More precisely, we have

Proposition 2.4 (Mass conservation).

For every n+n\in\mathbb{Z}^{+}, it holds ϕn¯=ϕ0¯\overline{\phi^{n}}=\overline{\phi^{0}}. As a consequence, we have <ϕnϕ0¯,1>Ω=0\big{<}\phi^{n}-\overline{\phi^{0}},1\big{>}_{\Omega}=0 for all n+n\in\mathbb{Z}^{+}.

Remark 2.3.

For the Cahn-Hilliard-Willmore system, namely, the case with η0\eta\leq 0 and λ>0\lambda>0, using a similar idea for (2.18)–(2.19), we can propose the following convex-splitting scheme:

ϕn+1ϕnΔt=Δhμn+1,\displaystyle\frac{\phi^{n+1}-\phi^{n}}{\Delta t}=\Delta_{h}\mu^{n+1}, (2.20)
μn+1=ϵ4Δh2ϕn+1+β(ϕn+1)β(ϕn+1)+ϵ2ζ=x,y,zβ′′(ϕn+1)aζ(|Dζϕn+1|2)\displaystyle\mu^{n+1}=\epsilon^{4}\Delta_{h}^{2}\phi^{n+1}+\beta(\phi^{n+1})\beta^{\prime}(\phi^{n+1})+\epsilon^{2}\sum_{\zeta=x,y,z}\beta^{\prime\prime}(\phi^{n+1})a_{\zeta}\left(|D_{\zeta}\phi^{n+1}|^{2}\right)
2ϵ2ζ=x,y,zdζ(Aζ(β(ϕn+1))Dζϕn+1)+λ2ϕn+1+ϵp+2ηΔhϕn+1\displaystyle\qquad\quad-2\epsilon^{2}\sum_{\zeta=x,y,z}d_{\zeta}\left(A_{\zeta}\big{(}\beta^{\prime}(\phi^{n+1})\big{)}D_{\zeta}\phi^{n+1}\right)+\lambda^{2}\phi^{n+1}+\epsilon^{p+2}\eta\Delta_{h}\phi^{n+1}
ϵpηβ(ϕn+1)+λϵpηϕn+2λϵ2Δhϕnλϕnβ(ϕn)λβ(ϕn).\displaystyle\qquad\quad-\epsilon^{p}\eta\beta(\phi^{n+1})+\lambda\epsilon^{p}\eta\phi^{n}+2\lambda\epsilon^{2}\Delta_{h}\phi^{n}-\lambda\phi^{n}\beta^{\prime}(\phi^{n})-\lambda\beta(\phi^{n}). (2.21)

Furthermore, when η0\eta\leq 0 and λ0\lambda\leq 0, we can replace the last four terms on the right-hand side of (2.21) by λϵpηϕn+1+2λϵ2Δhϕn+1λϕn+1β(ϕn+1)λβ(ϕn+1)\lambda\epsilon^{p}\eta\phi^{n+1}+2\lambda\epsilon^{2}\Delta_{h}\phi^{n+1}-\lambda\phi^{n+1}\beta^{\prime}(\phi^{n+1})-\lambda\beta(\phi^{n+1}). Theoretic analysis and numerical experiments of these modified schemes will be carried out in the future study.

3 Unique Solvability and Positivity-Preserving Property

Our aim in this section is to show that for every nn\in\mathbb{N}, the numerical scheme (2.18)–(2.19) is uniquely solvable for any time step size Δt>0\Delta t>0 and grid size h>0h>0. In particular, the discrete solution ϕn\phi^{n} satisfied the positivity-preserving property such that it is strictly separated from the pure states ±1\pm 1, i.e., 1<ϕi,j,kn<1-1<\phi^{n}_{i,j,k}<1 (cf. [28, 29] for the classical Cahn-Hilliard equation).

The main result of this section reads as follows:

Theorem 3.1 (Unique solvability and positivity-preserving property).

Let nn\in\mathbb{N} and Δt,h>0\Delta t,h>0. Given ϕn𝒞per\phi^{n}\in\mathcal{C}_{\mathrm{per}} with ϕn<1\|\phi^{n}\|_{\infty}<1, then the discrete system (2.18)–(2.19) admits a unique solution ϕn+1𝒞per\phi^{n+1}\in\mathcal{C}_{\mathrm{per}} such that

ϕn+1ϕ0¯𝒞̊perandϕn+1<1.\displaystyle\phi^{n+1}-\overline{\phi^{0}}\in\mathring{\mathcal{C}}_{\mathrm{per}}\quad\text{and}\quad\|\phi^{n+1}\|_{\infty}<1. (3.1)
Proof.

We divide the proof of Theorem 3.1 into several steps.

Step 1. Define the discrete energy:

𝒥n(ϕ)\displaystyle\mathcal{J}^{n}(\phi) =12Δtϕϕn1,h2+Ehc(ϕ)+fn,ϕΩ,\displaystyle=\frac{1}{2\Delta t}\|\phi-\phi^{n}\|_{-1,h}^{2}+E_{h}^{c}(\phi)+\left<f^{n},\phi\right>_{\Omega}, (3.2)

where Ehc(ϕ)E_{h}^{c}(\phi) is given by (2.16) and

fn\displaystyle f^{n} :=ϵ2(2λ+ϵpη)Δhϕnλϕnβ(ϕn)(λ+ϵpη)β(ϕn).\displaystyle:=\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta_{h}\phi^{n}-\lambda\phi^{n}\beta^{\prime}(\phi^{n})-(\lambda+\epsilon^{p}\eta)\beta(\phi^{n}). (3.3)

In view of Proposition 2.4, we introduce the admissible set

𝒜h:={ϕ𝒞per|ϕ<1,ϕϕ0¯,1Ω=0}N3.\mathcal{A}_{h}:=\left\{\phi\in\mathcal{C}_{\mathrm{per}}\,\left|\ \|\phi\|_{\infty}<1,\ \left<\phi-\overline{\phi^{0}},1\right>_{\Omega}=0\right.\right\}\subset\mathbb{R}^{N^{3}}. (3.4)

The existence of a solution to the discrete system (2.18)–(2.19) can be proved by seeking a minimizer of the energy 𝒥n\mathcal{J}^{n} over the set 𝒜h\mathcal{A}_{h}. For the sake of convenience and thanks to the mass conservation, we introduce the new variable with zero mean

φ=ϕϕ0¯\varphi=\phi-\overline{\phi^{0}}

and rewrite 𝒥n\mathcal{J}^{n} into the following equivalent form:

n(φ)\displaystyle\mathcal{F}^{n}(\varphi) :=𝒥n(φ+ϕ0¯)\displaystyle:=\mathcal{J}^{n}(\varphi+\overline{\phi^{0}})
=12Δtφ+ϕ0¯ϕn1,h2+ϵ42Δhφ22+12β(φ+ϕ0¯)22\displaystyle=\frac{1}{2\Delta t}\|\varphi+\overline{\phi^{0}}-\phi^{n}\|_{-1,h}^{2}+\frac{\epsilon^{4}}{2}\|\Delta_{h}\varphi\|_{2}^{2}+\frac{1}{2}\|\beta(\varphi+\overline{\phi^{0}})\|_{2}^{2}
+ϵ2β(φ+ϕ0¯),ζ=x,y,zaζ(|Dζϕn+1|2)Ω+λ(λ+ϵpη)2φ+ϕ0¯22\displaystyle\quad+\epsilon^{2}\left<\beta^{\prime}(\varphi+\overline{\phi^{0}}),\sum_{\zeta=x,y,z}a_{\zeta}\left(\left|D_{\zeta}\phi^{n+1}\right|^{2}\right)\right>_{\Omega}+\frac{\lambda(\lambda+\epsilon^{p}\eta)}{2}\|\varphi+\overline{\phi^{0}}\|_{2}^{2}
+fn,φ+ϕ0¯Ω.\displaystyle\quad+\left<f^{n},\varphi+\overline{\phi^{0}}\right>_{\Omega}.

Then n\mathcal{F}^{n} is defined on the shifted admissible set given by

𝒜̊h:={φ𝒞̊per|1ϕ0¯<φi,j,k<1ϕ0¯,i,j,k}N3.\mathring{\mathcal{A}}_{h}:=\left\{\varphi\in\mathring{\mathcal{C}}_{\mathrm{per}}\,\Big{|}\,-1-\overline{\phi^{0}}<\varphi_{i,j,k}<1-\overline{\phi^{0}},\ \forall i,j,k\right\}\subset\mathbb{R}^{N^{3}}.

It is straightforward to check that if φ𝒜̊h\varphi\in\mathring{\mathcal{A}}_{h} minimizes n\mathcal{F}^{n}, then ϕ:=φ+ϕ0¯𝒜h\phi:=\varphi+\overline{\phi^{0}}\in\mathcal{A}_{h} minimizes 𝒥n\mathcal{J}^{n}, and vice versa.

Step 2. To show the existence of a minimizer of n\mathcal{F}^{n} in 𝒜̊h\mathring{\mathcal{A}}_{h}, for any given δ(0,1/2)\delta\in(0,1/2), let us consider the auxiliary set

𝒜̊h,δ:={φ𝒞̊per|1ϕ0¯+δφ1ϕ0¯δ}𝒜̊h,\mathring{\mathcal{A}}_{h,\delta}:=\left\{\varphi\in\mathring{\mathcal{C}}_{\mathrm{per}}\,\Big{|}\,-1-\overline{\phi^{0}}+\delta\leq\varphi\leq 1-\overline{\phi^{0}}-\delta\right\}\subset\mathring{\mathcal{A}}_{h}, (3.5)

which is a bounded, closed and convex subset of 𝒞̊per\mathring{\mathcal{C}}_{\mathrm{per}}. Thus from the construction of n\mathcal{F}^{n}, we easily infer that it admits at least one minimizer over 𝒜̊h,δ\mathring{\mathcal{A}}_{h,\delta}.

Step 3. We now derive the key property that every minimizer of n\mathcal{F}^{n} cannot occur on the boundary of 𝒜̊h,δ\mathring{\mathcal{A}}_{h,\delta}, provided that the parameter δ\delta is sufficiently small. Here, by the boundary of 𝒜̊h,δ\mathring{\mathcal{A}}_{h,\delta}, we mean the set of functions ψ𝒜̊h,δ\psi\in\mathring{\mathcal{A}}_{h,\delta} such that ψ+ϕ0¯=1δ\|\psi+\overline{\phi^{0}}\|_{\infty}=1-\delta.

The above claim can be proved by using a contradiction argument (cf. [29] for the Cahn-Hilliard equation). Suppose that for some δ(0,1/2)\delta\in(0,1/2), a minimizer of n\mathcal{F}^{n} in 𝒜̊h,δ\mathring{\mathcal{A}}_{h,\delta}, which is denoted by φ\varphi^{*}, occurs at a boundary point of 𝒜̊h,δ\mathring{\mathcal{A}}_{h,\delta}. Then there is at least one grid point α0=(i0,j0,k0)\vec{\alpha}_{0}=(i_{0},j_{0},k_{0}) such that |φα0+ϕ0¯|=1δ|\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}}|=1-\delta. Without loss of generality, we assume that φα0+ϕ0¯=1+δ\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}}=-1+\delta. This implies that the grid function φ\varphi^{*} takes its (global) minimum at α0\vec{\alpha}_{0}. On the other hand, let α1=(i1,j1,k1)\vec{\alpha}_{1}=(i_{1},j_{1},k_{1}) be a grid point at which φ\varphi^{*} achieves its maximum. It follows that φα1+ϕ0¯1δ\varphi_{\vec{\alpha}_{1}}^{*}+\overline{\phi^{0}}\leq 1-\delta, since φ𝒜̊h,δ\varphi^{*}\in\mathring{\mathcal{A}}_{h,\delta}. Besides, from the fact φ¯=0\overline{\varphi^{*}}=0, we infer that φα10\varphi_{\vec{\alpha}_{1}}^{*}\geq 0.

Since n\mathcal{F}^{n} is indeed smooth over 𝒜̊h,δ\mathring{\mathcal{A}}_{h,\delta}, then its directional derivative can be calculated as

dn(φ+sψ)ds|s=0\displaystyle\frac{\mathrm{d}\mathcal{F}^{n}(\varphi^{*}+s\psi)}{\mathrm{d}s}\Big{|}_{s=0}
=1Δt(Δh)1(φϕn+ϕ0¯),ψΩ+ϵ4Δh2φ,ψ+λ(λ+ϵpη)φ+ϕ0¯,ψΩ\displaystyle\quad=\frac{1}{\Delta t}\left<(-\Delta_{h})^{-1}(\varphi^{*}-\phi^{n}+\overline{\phi^{0}}),\psi\right>_{\Omega}+\epsilon^{4}\left<\Delta_{h}^{2}\varphi^{*},\psi\right>+\lambda(\lambda+\epsilon^{p}\eta)\left<\varphi^{*}+\overline{\phi^{0}},\psi\right>_{\Omega}
+ϵ2ζ=x,y,z[β′′(φ+ϕ0¯)aζ(|Dζφ|2)2dζ(Aζ(β(φ+ϕ0¯))Dζφ)],ψΩ\displaystyle\qquad+\epsilon^{2}\left<\sum_{\zeta=x,y,z}\left[\beta^{\prime\prime}(\varphi^{*}+\overline{\phi^{0}})a_{\zeta}\left(|D_{\zeta}\varphi^{*}|^{2}\right)-2d_{\zeta}\left(A_{\zeta}\big{(}\beta^{\prime}(\varphi^{*}+\overline{\phi^{0}})\big{)}D_{\zeta}\varphi^{*}\right)\right],\psi\right>_{\Omega}
+β(φ+ϕ0¯)β(φ+ϕ0¯),ψΩ+fn,ψΩ,ψ𝒞̊per.\displaystyle\qquad+\left<\beta(\varphi^{*}+\overline{\phi^{0}})\beta^{\prime}(\varphi^{*}+\overline{\phi^{0}}),\psi\right>_{\Omega}+\left<f^{n},\psi\right>_{\Omega},\quad\forall\,\psi\in\mathring{\mathcal{C}}_{\mathrm{per}}.

In particular, we choose the direction ψ𝒞̊per\psi\in\mathring{\mathcal{C}}_{\mathrm{per}} with ψi,j,k=δi,i0δj,j0δk,k0δi,i1δj,j1δk,k1\psi_{i,j,k}=\delta_{i,i_{0}}\delta_{j,j_{0}}\delta_{k,k_{0}}-\delta_{i,i_{1}}\delta_{j,j_{1}}\delta_{k,k_{1}}. This yields that

1h3dn(φ+sψ)ds|s=0\displaystyle\frac{1}{h^{3}}\frac{\mathrm{d}\mathcal{F}^{n}(\varphi^{*}+s\psi)}{\mathrm{d}s}\Big{|}_{s=0}
=1Δt(Δh)1(φϕn+ϕ0¯)α01Δt(Δh)1(φϕn+ϕ0¯)α1\displaystyle\quad=\frac{1}{\Delta t}(-\Delta_{h})^{-1}(\varphi^{*}-\phi^{n}+\overline{\phi^{0}})_{\vec{\alpha}_{0}}-\frac{1}{\Delta t}(-\Delta_{h})^{-1}(\varphi^{*}-\phi^{n}+\overline{\phi^{0}})_{\vec{\alpha}_{1}}
+ϵ4Δh2(φα0φα1)+λ(λ+ϵpη)(φα0φα1)\displaystyle\qquad+\epsilon^{4}\Delta_{h}^{2}(\varphi_{\vec{\alpha}_{0}}^{*}-\varphi_{\vec{\alpha}_{1}}^{*})+\lambda(\lambda+\epsilon^{p}\eta)(\varphi_{\vec{\alpha}_{0}}^{*}-\varphi_{\vec{\alpha}_{1}}^{*})
+ϵ2(ζ=x,y,z[β′′(φα0+ϕ0¯)aζ(|Dζφα0|2)2dζ(Aζ(β(φα0+ϕ0¯))Dζφα0)])\displaystyle\qquad+\epsilon^{2}\left(\sum_{\zeta=x,y,z}\left[\beta^{\prime\prime}(\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}})a_{\zeta}\left(|D_{\zeta}\varphi_{\vec{\alpha}_{0}}^{*}|^{2}\right)-2d_{\zeta}\left(A_{\zeta}\big{(}\beta^{\prime}(\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}})\big{)}D_{\zeta}\varphi_{\vec{\alpha}_{0}}^{*}\right)\right]\right)
ϵ2(ζ=x,y,z[β′′(φα1+ϕ0¯)aζ(|Dζφα1|2)2dζ(Aζ(β(φα1+ϕ0¯))Dζφα1)])\displaystyle\qquad-\epsilon^{2}\left(\sum_{\zeta=x,y,z}\left[\beta^{\prime\prime}(\varphi_{\vec{\alpha}_{1}}^{*}+\overline{\phi^{0}})a_{\zeta}\left(|D_{\zeta}\varphi_{\vec{\alpha}_{1}}^{*}|^{2}\right)-2d_{\zeta}\left(A_{\zeta}\big{(}\beta^{\prime}(\varphi_{\vec{\alpha}_{1}}^{*}+\overline{\phi^{0}})\big{)}D_{\zeta}\varphi_{\vec{\alpha}_{1}}^{*}\right)\right]\right)
+β(φα0+ϕ0¯)β(φα0+ϕ0¯)β(φα1+ϕ0¯)β(φα1+ϕ0¯)\displaystyle\qquad+\beta(\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}})\beta^{\prime}(\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}})-\beta(\varphi_{\vec{\alpha}_{1}}^{*}+\overline{\phi^{0}})\beta^{\prime}(\varphi_{\vec{\alpha}_{1}}^{*}+\overline{\phi^{0}})
+(fα0nfα1n).\displaystyle\qquad+(f_{\vec{\alpha}_{0}}^{n}-f_{\vec{\alpha}_{1}}^{n}). (3.6)

The first two terms on the right-hand side of (3.6) can be estimated by using Lemma 2.2 such that

2C0(Δh)1(φϕn+ϕ0¯)α0(Δh)1(φϕn+ϕ0¯)α12C0.-2C_{0}\leq(-\Delta_{h})^{-1}(\varphi^{*}-\phi^{n}+\overline{\phi^{0}})_{\vec{\alpha}_{0}}-(-\Delta_{h})^{-1}(\varphi^{*}-\phi^{n}+\overline{\phi^{0}})_{\vec{\alpha}_{1}}\leq 2C_{0}. (3.7)

For the third and fourth terms, recalling the definition of Δh\Delta_{h} and (3.5), we find that

Δh2(φα0φα1)288(1δ)h4288h4,\displaystyle\|\Delta_{h}^{2}(\varphi_{\vec{\alpha}_{0}}^{*}-\varphi_{\vec{\alpha}_{1}}^{*})\|_{\infty}\leq\frac{288(1-\delta)}{h^{4}}\leq\frac{288}{h^{4}}, (3.8)
φα0φα12(1δ)2.\displaystyle\|\varphi_{\vec{\alpha}_{0}}^{*}-\varphi_{\vec{\alpha}_{1}}^{*}\|_{\infty}\leq 2(1-\delta)\leq 2. (3.9)

Estimate for the fifth term involving β′′(φα0+ϕ0¯)aζ(|Dζφα0|2)2dζ(Aζ(β(φα0+ϕ0¯))Dζφα0)\beta^{\prime\prime}(\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}})a_{\zeta}\big{(}|D_{\zeta}\varphi_{\vec{\alpha}_{0}}^{*}|^{2}\big{)}-2d_{\zeta}\big{(}A_{\zeta}\big{(}\beta^{\prime}(\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}})\big{)}D_{\zeta}\varphi_{\vec{\alpha}_{0}}^{*}\big{)} turns out to be more tricky. For simplicity, we only consider the estimate along the xx-direction. A direct calculation yields that

h2[β′′(φα0+ϕ0¯)ax(|Dxφα0|2)2dx(Ax(β(φα0+ϕ0¯))Dxφα0)]\displaystyle h^{2}\Big{[}\beta^{\prime\prime}(\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}})a_{x}\left(|D_{x}\varphi_{\vec{\alpha}_{0}}^{*}|^{2}\right)-2d_{x}\left(A_{x}\big{(}\beta^{\prime}(\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}})\big{)}D_{x}\varphi_{\vec{\alpha}_{0}}^{*}\right)\Big{]}
=12β′′(φi0+ϕ0¯)((φi0+1φi0)2+(φi0φi01)2)\displaystyle\quad=\frac{1}{2}\beta^{\prime\prime}(\varphi_{i_{0}}^{*}+\overline{\phi^{0}})\left((\varphi_{i_{0}+1}^{*}-\varphi_{i_{0}}^{*})^{2}+(\varphi_{i_{0}}^{*}-\varphi_{i_{0}-1}^{*})^{2}\right)
(β(φi0+ϕ0¯)+β(φi0+1+ϕ0¯))(φi0+1φi0)\displaystyle\qquad-\left(\beta^{\prime}(\varphi_{i_{0}}^{*}+\overline{\phi^{0}})+\beta^{\prime}(\varphi_{i_{0}+1}^{*}+\overline{\phi^{0}})\right)(\varphi_{i_{0}+1}^{*}-\varphi_{i_{0}}^{*})
+(β(φi0+ϕ0¯)+β(φi01+ϕ0¯))(φi0φi01).\displaystyle\qquad+\left(\beta^{\prime}(\varphi_{i_{0}}^{*}+\overline{\phi^{0}})+\beta^{\prime}(\varphi_{i_{0}-1}^{*}+\overline{\phi^{0}})\right)(\varphi_{i_{0}}^{*}-\varphi_{i_{0}-1}^{*}).

Here, we denote φi,j0,k0\varphi_{i,j_{0},k_{0}}^{*} by φi\varphi_{i}^{*} for simplicity (as we only consider the xx-direction). Using Taylor’s expansion for β(φi0+1+ϕ0¯)\beta^{\prime}(\varphi_{i_{0}+1}^{*}+\overline{\phi^{0}}) and β(φi01+ϕ0¯)\beta^{\prime}(\varphi_{i_{0}-1}^{*}+\overline{\phi^{0}}) at φi0+ϕ0¯\varphi_{i_{0}}^{*}+\overline{\phi^{0}}, we obtain

h2[β′′(φα0+ϕ0¯)ax(|Dxφα0|2)2dx(Ax(β(φα0+ϕ0¯))Dxφα0)]\displaystyle h^{2}\Big{[}\beta^{\prime\prime}(\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}})a_{x}\left(|D_{x}\varphi_{\vec{\alpha}_{0}}^{*}|^{2}\right)-2d_{x}\left(A_{x}\big{(}\beta^{\prime}(\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}})\big{)}D_{x}\varphi_{\vec{\alpha}_{0}}^{*}\right)\Big{]}
=12β′′(φi0+ϕ0¯)((φi0+1φi0)2+(φi0φi01)2)\displaystyle\quad=\frac{1}{2}\beta^{\prime\prime}(\varphi_{i_{0}}^{*}+\overline{\phi^{0}})\left((\varphi_{i_{0}+1}^{*}-\varphi_{i_{0}}^{*})^{2}+(\varphi_{i_{0}}^{*}-\varphi_{i_{0}-1}^{*})^{2}\right)
(β(φi0+ϕ0¯)+12β(φi0+1+ϕ0¯))(φi0+1φi0)\displaystyle\qquad-\left(\beta^{\prime}(\varphi_{i_{0}}^{*}+\overline{\phi^{0}})+\frac{1}{2}\beta^{\prime}(\varphi_{i_{0}+1}^{*}+\overline{\phi^{0}})\right)(\varphi_{i_{0}+1}^{*}-\varphi_{i_{0}}^{*})
+(β(φi0+ϕ0¯)+12β(φi01+ϕ0¯))(φi0φi01)\displaystyle\qquad+\left(\beta^{\prime}(\varphi_{i_{0}}^{*}+\overline{\phi^{0}})+\frac{1}{2}\beta^{\prime}(\varphi_{i_{0}-1}^{*}+\overline{\phi^{0}})\right)(\varphi_{i_{0}}^{*}-\varphi_{i_{0}-1}^{*})
12β(φi0+ϕ0¯)(φi0+1φi0)12β′′(φi0+ϕ0¯)(φi0+1φi0)2β′′′(ξ1+ϕ0¯)4(φi0+1φi0)3\displaystyle\qquad-\frac{1}{2}\beta^{\prime}(\varphi_{i_{0}}^{*}+\overline{\phi^{0}})(\varphi_{i_{0}+1}^{*}-\varphi_{i_{0}}^{*})-\frac{1}{2}\beta^{\prime\prime}(\varphi_{i_{0}}^{*}+\overline{\phi^{0}})(\varphi_{i_{0}+1}^{*}-\varphi_{i_{0}}^{*})^{2}-\frac{\beta^{\prime\prime\prime}(\xi_{1}+\overline{\phi^{0}})}{4}(\varphi_{i_{0}+1}^{*}-\varphi_{i_{0}}^{*})^{3}
+12β(φi0+ϕ0¯)(φi0φi01)12β′′(φi0+ϕ0¯)(φi0φi01)2+β′′′(ξ2+ϕ0¯)4(φi0φi01)3\displaystyle\qquad+\frac{1}{2}\beta^{\prime}(\varphi_{i_{0}}^{*}+\overline{\phi^{0}})(\varphi_{i_{0}}^{*}-\varphi_{i_{0}-1}^{*})-\frac{1}{2}\beta^{\prime\prime}(\varphi_{i_{0}}^{*}+\overline{\phi^{0}})(\varphi_{i_{0}}^{*}-\varphi_{i_{0}-1}^{*})^{2}+\frac{\beta^{\prime\prime\prime}(\xi_{2}+\overline{\phi^{0}})}{4}(\varphi_{i_{0}}^{*}-\varphi_{i_{0}-1}^{*})^{3}
=(32β(φi0+ϕ0¯)+12β(φi0+1+ϕ0¯))(φi0+1φi0)\displaystyle\quad=-\left(\frac{3}{2}\beta^{\prime}(\varphi_{i_{0}}^{*}+\overline{\phi^{0}})+\frac{1}{2}\beta^{\prime}(\varphi_{i_{0}+1}^{*}+\overline{\phi^{0}})\right)(\varphi_{i_{0}+1}^{*}-\varphi_{i_{0}}^{*})
+(32β(φi0+ϕ0¯)+12β(φi01+ϕ0¯))(φi0φi01)\displaystyle\qquad+\left(\frac{3}{2}\beta^{\prime}(\varphi_{i_{0}}^{*}+\overline{\phi^{0}})+\frac{1}{2}\beta^{\prime}(\varphi_{i_{0}-1}^{*}+\overline{\phi^{0}})\right)(\varphi_{i_{0}}^{*}-\varphi_{i_{0}-1}^{*})
β′′′(ξ1+ϕ0¯)4(φi0+1φi0)3+β′′′(ξ2+ϕ0¯)4(φi0φi01)3\displaystyle\qquad-\frac{\beta^{\prime\prime\prime}(\xi_{1}+\overline{\phi^{0}})}{4}(\varphi_{i_{0}+1}^{*}-\varphi_{i_{0}}^{*})^{3}+\frac{\beta^{\prime\prime\prime}(\xi_{2}+\overline{\phi^{0}})}{4}(\varphi_{i_{0}}^{*}-\varphi_{i_{0}-1}^{*})^{3}
0.\displaystyle\quad\leq 0. (3.10)

Here ξ1\xi_{1} is between φi0\varphi_{i_{0}}^{*} and φi0+1\varphi_{i_{0}+1}^{*}, while ξ2\xi_{2} is between φi0\varphi_{i_{0}}^{*} and φi01\varphi_{i_{0}-1}^{*}. The last inequality in (3.10) holds thanks to the definition of α0\vec{\alpha}_{0} (i.e., the minimum point of φ\varphi^{*}) so that φi0+1φi00\varphi_{i_{0}+1}^{*}-\varphi_{i_{0}}^{*}\geq 0, φi0φi010\varphi_{i_{0}}^{*}-\varphi_{i_{0}-1}^{*}\leq 0, and the fact that β(r),β′′′(r)>0\beta^{\prime}(r),\ \beta^{\prime\prime\prime}(r)>0 for any r(1,1)r\in(-1,1). Similar results can be derived for the yy, zz direction. Besides, for the term involving α1\vec{\alpha}_{1}, we can conclude

β′′(φα1+ϕ0¯)aζ(|Dζφα1|2)2dζ(Aζ(β(φα1+ϕ0¯))Dζφα1)0,forζ=x,y,z.\beta^{\prime\prime}(\varphi_{\vec{\alpha}_{1}}^{*}+\overline{\phi^{0}})a_{\zeta}\left(|D_{\zeta}\varphi_{\vec{\alpha}_{1}}^{*}|^{2}\right)-2d_{\zeta}\left(A_{\zeta}\big{(}\beta^{\prime}(\varphi_{\vec{\alpha}_{1}}^{*}+\overline{\phi^{0}})\big{)}D_{\zeta}\varphi_{\vec{\alpha}_{1}}^{*}\right)\geq 0,\quad\text{for}\ \zeta=x,y,z. (3.11)

Next, we infer from (2.10)–(2.11) that the nonlinear function ββ\beta\beta^{\prime} is strictly increasing on (1,1)(-1,1). Since φα0+ϕ0¯=1+δ\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}}=-1+\delta and φα10\varphi_{\vec{\alpha}_{1}}^{*}\geq 0, then we have

β(φα0+ϕ0¯)β(φα0+ϕ0¯)β(φα1+ϕ0¯)β(φα1+ϕ0¯)\displaystyle\beta(\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}})\beta^{\prime}(\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}})-\beta(\varphi_{\vec{\alpha}_{1}}^{*}+\overline{\phi^{0}})\beta^{\prime}(\varphi_{\vec{\alpha}_{1}}^{*}+\overline{\phi^{0}})
(1δ+12δ)lnδ2δ(11+ϕ0¯+11ϕ0¯)ln1+ϕ0¯1ϕ0¯.\displaystyle\quad\leq\left(\frac{1}{\delta}+\frac{1}{2-\delta}\right)\ln\frac{\delta}{2-\delta}-\left(\frac{1}{1+\overline{\phi^{0}}}+\frac{1}{1-\overline{\phi^{0}}}\right)\ln\frac{1+\overline{\phi^{0}}}{1-\overline{\phi^{0}}}. (3.12)

Finally, we treat the term involving fnf^{n} (see (3.3)). From the assumption ϕn<1\|\phi^{n}\|_{\infty}<1 and the fact that for any given h>0h>0, there are only finite number of grid points, we can find some δn(0,1)\delta_{n}\in(0,1) such that

1+δnϕi,j,kn1δn, 1i,j,kN.-1+\delta_{n}\leq\phi_{i,j,k}^{n}\leq 1-\delta_{n},\quad\forall\,1\leq i,j,k\leq N. (3.13)

Here, the constant δn\delta_{n} may depend on the parameter nn. From the refined bound (3.13) and the assumption λ,η,ϵ>0\lambda,\eta,\epsilon>0, we see that

|fn|24ϵ2(2λ+ϵpη)(1δn)h22λ(1+δn)(1δn+12δn)2(λ+ϵpη)lnδn2δn.|f^{n}|\leq\frac{24\epsilon^{2}(2\lambda+\epsilon^{p}\eta)(1-\delta_{n})}{h^{2}}-2\lambda(-1+\delta_{n})\left(\frac{1}{\delta_{n}}+\frac{1}{2-\delta_{n}}\right)-2(\lambda+\epsilon^{p}\eta)\ln\frac{\delta_{n}}{2-\delta_{n}}. (3.14)

Collecting the estimates (3.7)–(3.14), we can deduce from (3.6) that

1h3dn(φ+sψ)ds|s=0g(δ)+max{C^,1},\displaystyle\frac{1}{h^{3}}\frac{\mathrm{d}\mathcal{F}^{n}(\varphi^{*}+s\psi)}{\mathrm{d}s}\Big{|}_{s=0}\leq g(\delta)+\max\{\widehat{C},1\},

where

g(δ)=(1δ+12δ)lnδ2δg(\delta)=\left(\frac{1}{\delta}+\frac{1}{2-\delta}\right)\ln\frac{\delta}{2-\delta}

and

C^\displaystyle\widehat{C} =2C2Δt+288ϵ4h4+2λ(λ+ϵpη)(11+ϕ0¯+11ϕ0¯)ln1+ϕ0¯1ϕ0¯\displaystyle=\frac{2C_{2}}{\Delta t}+\frac{288\epsilon^{4}}{h^{4}}+2\lambda(\lambda+\epsilon^{p}\eta)-\left(\frac{1}{1+\overline{\phi^{0}}}+\frac{1}{1-\overline{\phi^{0}}}\right)\ln\frac{1+\overline{\phi^{0}}}{1-\overline{\phi^{0}}}
+24ϵ2(2λ+ϵpη)(1δn)h22λ(1+δn)(1δn+12δn)\displaystyle\quad+\frac{24\epsilon^{2}(2\lambda+\epsilon^{p}\eta)(1-\delta_{n})}{h^{2}}-2\lambda(-1+\delta_{n})\left(\frac{1}{\delta_{n}}+\frac{1}{2-\delta_{n}}\right)
2(λ+ϵpη)lnδn2δn.\displaystyle\quad-2(\lambda+\epsilon^{p}\eta)\ln\frac{\delta_{n}}{2-\delta_{n}}.

Observe that C^\widehat{C} is a constant that may depend on Δt,h,δn\Delta t,\ h,\ \delta_{n}, ϕ0¯\overline{\phi^{0}}, λ\lambda, η\eta, ϵ\epsilon, but is independent of δ\delta. Since

limδ0+g(δ)=,\lim_{\delta\searrow 0^{+}}g(\delta)=-\infty,

and g(δ)=β(1+δ)β(1+δ)g(\delta)=\beta(-1+\delta)\beta^{\prime}(-1+\delta) is strictly increasing on (0,1)(0,1), we can find some δ^(0,1/2)\widehat{\delta}\in(0,1/2) sufficiently small such that

g(δ)+max{C^,1}1,δ(0,δ^].\displaystyle g(\delta)+\max\{\widehat{C},1\}\leq-1,\quad\forall\,\delta\in(0,\widehat{\delta}\,]. (3.15)

which gives

dn(φ+sψ)ds|s=0<0.\frac{\mathrm{d}\mathcal{F}^{n}(\varphi^{*}+s\psi)}{\mathrm{d}s}\Big{|}_{s=0}<0.

This contradicts the assumption that n\mathcal{F}^{n} attains its minimum at φ\varphi^{*}, because the directional derivative is strictly negative along a specific direction pointing into the interior of 𝒜̊h,δ\mathring{\mathcal{A}}_{h,\delta}.

By a similar argument, we can show that for any δ(0,δ^]\delta\in(0,\widehat{\delta}\,], a minimizer of n\mathcal{F}^{n} over 𝒜̊h,δ\mathring{\mathcal{A}}_{h,\delta} cannot occur at a boundary point φ\varphi^{*} such that φα0+ϕ0¯=1δ\varphi_{\vec{\alpha}_{0}}^{*}+\overline{\phi^{0}}=1-\delta, for some α0\vec{\alpha}_{0}.

In summary, there exists a δ^(0,1/2)\widehat{\delta}\in(0,1/2) such that for every δ(0,δ^]\delta\in(0,\widehat{\delta}\,], the (global) minimum of n\mathcal{F}^{n} over the set 𝒜̊h,δ\mathring{\mathcal{A}}_{h,\delta} exists and it can be only attained at a certain interior point φ𝒜̊h,δ\varphi^{*}\in\mathring{\mathcal{A}}_{h,\delta}.

Step 4. Let φ\varphi^{*} be a minimizer of n\mathcal{F}^{n} over 𝒜̊h,δ^\mathring{\mathcal{A}}_{h,\widehat{\delta}}, where δ^\widehat{\delta} is given in the previous step (see (3.15)). Below we show that φ\varphi^{*} must be the unique minimizer of n\mathcal{F}^{n} over 𝒜̊h\mathring{\mathcal{A}}_{h}. In other words, ϕ=φ+ϕ0¯\phi^{*}=\varphi^{*}+\overline{\phi^{0}} is the unique minimizer of 𝒥n\mathcal{J}^{n} over 𝒜h\mathcal{A}_{h}.

First, for every δ(0,δ^]\delta\in(0,\widehat{\delta}\,], if φ\varphi^{*} is a minimizer of n\mathcal{F}^{n} over 𝒜̊h,δ\mathring{\mathcal{A}}_{h,\delta}, then it is unique. This is equivalent to say that ϕ=φ+ϕ0¯\phi^{*}=\varphi^{*}+\overline{\phi^{0}} is the unique minimizer of 𝒥n\mathcal{J}^{n} over 𝒜h,δ\mathcal{A}_{h,\delta}. Indeed, similar to the proof of Proposition 2.3, we know that the discrete energy EhcE_{h}^{c} is strictly convex on 𝒜h,δ\mathcal{A}_{h,\delta}. Since the first term in 𝒥n\mathcal{J}^{n} is quadratic and the third term is linear in ϕ\phi, we deduce that 𝒥n\mathcal{J}^{n} is strictly convex on the set 𝒜h,δ\mathcal{A}_{h,\delta} as well. Thanks to the strict convexity of 𝒥n\mathcal{J}^{n} over 𝒜h,δ\mathcal{A}_{h,\delta}, the minimizer ϕ\phi^{*} is unique.

Next, for every δ(0,δ^)\delta\in(0,\widehat{\delta}\,), we denote the unique minimizer of n\mathcal{F}^{n} over 𝒜̊h,δ\mathring{\mathcal{A}}_{h,\delta} by φδ\varphi^{\delta}. Since 𝒜̊h,δ^𝒜̊h,δ\mathring{\mathcal{A}}_{h,\widehat{\delta}}\subset\mathring{\mathcal{A}}_{h,\delta}, then we have n(φδ)n(φ)\mathcal{F}^{n}(\varphi^{\delta})\leq\mathcal{F}^{n}(\varphi^{*}). Let us consider two cases.

Case 1. For all δ(0,δ^)\delta\in(0,\widehat{\delta}\,), it holds n(φδ)=n(φ)\mathcal{F}^{n}(\varphi^{\delta})=\mathcal{F}^{n}(\varphi^{*}). In this case, φ\varphi^{*} is a minimizer of n\mathcal{F}^{n} over 𝒜̊h,δ\mathring{\mathcal{A}}_{h,\delta} as well. Thanks to the uniqueness, it easily follows that φδ=φ\varphi^{\delta}=\varphi^{*} for any δ(0,δ^)\delta\in(0,\widehat{\delta}\,) and thus φ\varphi^{*} is indeed the unique minimizer of n\mathcal{F}^{n} over 𝒜̊h\mathring{\mathcal{A}}_{h}.

Case 2. There exists some δ~(0,δ^)\widetilde{\delta}\in(0,\widehat{\delta}\,), the corresponding unique minimizer of n\mathcal{F}^{n} over 𝒜̊h,δ~\mathring{\mathcal{A}}_{h,\widetilde{\delta}}, denoted by φ\varphi^{**}, satisfies n(φ)<n(φ)\mathcal{F}^{n}(\varphi^{**})<\mathcal{F}^{n}(\varphi^{*}). In this case, from Step 3, we can find a constant δ>δ~\delta^{**}>\widetilde{\delta}, such that φ+ϕ0¯=1δ\|\varphi^{**}+\overline{\phi^{0}}\|_{\infty}=1-\delta^{**}. Besides, since n(φ)<n(φ)\mathcal{F}^{n}(\varphi^{**})<\mathcal{F}^{n}(\varphi^{*}), we must have δ<δ^\delta^{**}<\widehat{\delta}. Now let us denote the unique minimizer of n\mathcal{F}^{n} over 𝒜̊h,δ\mathring{\mathcal{A}}_{h,\delta^{**}} by φ\varphi^{\sharp}. From Step 3, we find φ+ϕ0¯<1δ\|\varphi^{\sharp}+\overline{\phi^{0}}\|_{\infty}<1-\delta^{**}, which implies that φφ\varphi^{\sharp}\neq\varphi^{**}. On one hand, since φ𝒜̊h,δ\varphi^{**}\in\mathring{\mathcal{A}}_{h,\delta^{**}}, then n(φ)n(φ)\mathcal{F}^{n}(\varphi^{\sharp})\leq\mathcal{F}^{n}(\varphi^{**}). On the other hand, it follows from 𝒜̊h,δ𝒜̊h,δ~\mathring{\mathcal{A}}_{h,\delta^{**}}\subset\mathring{\mathcal{A}}_{h,\widetilde{\delta}} that n(φ)n(φ)\mathcal{F}^{n}(\varphi^{**})\leq\mathcal{F}^{n}(\varphi^{\sharp}). Thus, n(φ)=n(φ)\mathcal{F}^{n}(\varphi^{\sharp})=\mathcal{F}^{n}(\varphi^{**}) so that φ\varphi^{\sharp} should also be a minimizer of n\mathcal{F}^{n} over 𝒜̊h,δ~\mathring{\mathcal{A}}_{h,\widetilde{\delta}}. Thanks to the uniqueness of minimizer in 𝒜̊h,δ~\mathring{\mathcal{A}}_{h,\widetilde{\delta}}, we have φ=φ\varphi^{\sharp}=\varphi^{**}, which leads to a contradiction. Hence, the situation described in Case 2 will never happen.

In summary, the energy functional 𝒥n\mathcal{J}^{n} admits a unique minimizer ϕ\phi^{*} over the set 𝒜h\mathcal{A}_{h} such that φ=ϕϕ0¯\varphi^{*}=\phi^{*}-\overline{\phi^{0}} satisfies

dn(φ+sψ)ds|s=0=0,ψ𝒞̊per.\frac{\mathrm{d}\mathcal{F}^{n}(\varphi^{*}+s\psi)}{\mathrm{d}s}\Big{|}_{s=0}=0,\quad\forall\,\psi\in\mathring{\mathcal{C}}_{\mathrm{per}}.

This yields that ϕn+1=ϕ\phi^{n+1}=\phi^{*} is the unique solution to the discrete system (2.18)–(2.19). In particular, by the same argument as that for (3.13), we can find some δn+1(0,1/2)\delta_{n+1}\in(0,1/2) such that

1+δn+1ϕi,j,kn+11δn+1, 1i,j,kN.-1+\delta_{n+1}\leq\phi_{i,j,k}^{n+1}\leq 1-\delta_{n+1},\quad\forall\,1\leq i,j,k\leq N. (3.16)

The proof of Theorem 3.1 is complete. ∎

Remark 3.1.

In view of (3.15), the distance δn+1\delta_{n+1} between the discrete solution ϕn+1\phi^{n+1} and the pure states ±1\pm 1 may not be uniform with respect to the index nn. Nevertheless, when the spatial dimension is less than or equal to two, from Proposition 2.2, the convergence analysis and error estimates given in Section 5, we can derive a uniform separation property for the discrete solution ϕn+1\phi^{n+1} (see (5.31) below).

4 Unconditional Energy Stability

In this section, we show that the numerical scheme (2.18)–(2.19) is unconditionally energy stable. This property follows from the convex-concave decomposition of the discrete energy (recall Corollary 2.1).

Theorem 4.2 (Unconditional energy stability).

The discrete system (2.18)–(2.19) is unconditionally energy stable. Namely, for any nn\in\mathbb{N} and Δt>0\Delta t>0, we have

Eh(ϕn+1)+Δthμn+122Eh(ϕn),E_{h}(\phi^{n+1})+\Delta t\,\|\nabla_{h}\mu^{n+1}\|_{2}^{2}\leq E_{h}(\phi^{n}), (4.1)

where the discrete energy EhE_{h} is given by (2.15).

Proof.

Similar to the proof of Proposition 2.3, for the decomposed discrete energy Eh(ϕ)=Ehc(ϕ)Ehe(ϕ)E_{h}(\phi)=E_{h}^{c}(\phi)-E_{h}^{e}(\phi), we have

δϕEhc(ϕ)\displaystyle\delta_{\phi}E_{h}^{c}(\phi) =ϵ4Δh2ϕ+β(ϕ)β(ϕ)+ϵ2ζ=x,y,zβ′′(ϕ)aζ(|Dζϕ|2)\displaystyle=\epsilon^{4}\Delta_{h}^{2}\phi+\beta(\phi)\beta^{\prime}(\phi)+\epsilon^{2}\sum_{\zeta=x,y,z}\beta^{\prime\prime}(\phi)a_{\zeta}\left(|D_{\zeta}\phi|^{2}\right)
2ϵ2ζ=x,y,zdζ(Aζ(β(ϕ))Dζϕ)+λ(λ+ϵpη)ϕ,\displaystyle\quad-2\epsilon^{2}\sum_{\zeta=x,y,z}d_{\zeta}\left(A_{\zeta}\big{(}\beta^{\prime}(\phi)\big{)}D_{\zeta}\phi\right)+\lambda(\lambda+\epsilon^{p}\eta)\phi,
δϕEhe(ϕ)\displaystyle\delta_{\phi}E_{h}^{e}(\phi) =ϵ2(2λ+ϵpη)Δhϕn+λϕnβ(ϕn)+(λ+ϵpη)β(ϕn),\displaystyle=-\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta_{h}\phi^{n}+\lambda\phi^{n}\beta^{\prime}(\phi^{n})+(\lambda+\epsilon^{p}\eta)\beta(\phi^{n}),

for any ϕ𝒞per\phi\in\mathcal{C}_{\mathrm{per}} satisfying ϕ<1\|\phi\|_{\infty}<1. Thanks to Theorem 3.1 and Corollary 2.1, we can deduce the following inequalities for the discrete solution ϕn+1\phi^{n+1}:

Ehc(ϕn)Ehc(ϕn+1)δϕEhc(ϕn+1),ϕnϕn+1Ω,\displaystyle E_{h}^{c}(\phi^{n})-E_{h}^{c}(\phi^{n+1})\geq\left<\delta_{\phi}E_{h}^{c}(\phi^{n+1}),\phi^{n}-\phi^{n+1}\right>_{\Omega},
Ehe(ϕn+1)Ehe(ϕn)δϕEhe(ϕn),ϕn+1ϕnΩ.\displaystyle E_{h}^{e}(\phi^{n+1})-E_{h}^{e}(\phi^{n})\geq\left<\delta_{\phi}E_{h}^{e}(\phi^{n}),\phi^{n+1}-\phi^{n}\right>_{\Omega}.

Combining the above results with (2.18)–(2.19), and using the summation-by-parts formulae (see Lemma 2.1), we obtain the energy dissipative property (4.1):

Eh(ϕn+1)Eh(ϕn)\displaystyle E_{h}(\phi^{n+1})-E_{h}(\phi^{n}) =Ehc(ϕn+1)Ehc(ϕn)(Ehe(ϕn+1)Ehe(ϕn))\displaystyle=E_{h}^{c}(\phi^{n+1})-E_{h}^{c}(\phi^{n})-(E_{h}^{e}(\phi^{n+1})-E_{h}^{e}(\phi^{n}))
δϕEhc(ϕn+1)δϕEhe(ϕn),ϕn+1ϕnΩ\displaystyle\leq\left<\delta_{\phi}E_{h}^{c}(\phi^{n+1})-\delta_{\phi}E_{h}^{e}(\phi^{n}),\phi^{n+1}-\phi^{n}\right>_{\Omega}
=Δtμn+1,Δhμn+1Ω\displaystyle=\Delta t\left<\mu^{n+1},\Delta_{h}\mu^{n+1}\right>_{\Omega}
=Δthμn+1220.\displaystyle=-\Delta t\,\|\nabla_{h}\mu^{n+1}\|_{2}^{2}\leq 0.

The proof is complete. ∎

The energy dissipation property enables us to derive an uniform Hh2H_{h}^{2}-estimate for the solution to the discrete system (2.18)–(2.19).

Corollary 4.1 (Uniform Hh2H_{h}^{2}-estimate).

For every positive integer mm, the discrete solution ϕm\phi^{m} to the system (2.18)–(2.19) satisfies the following estimate

ϕmHh2C,\|\phi^{m}\|_{H^{2}_{h}}\leq C, (4.2)

where the constant C>0C>0 only depends on λ\lambda, η\eta, ϵ\epsilon, |Ω||\Omega| and ϕ0\phi^{0}.

Proof.

It follows from (4.1) that Eh(ϕm)Eh(ϕ0)E_{h}(\phi^{m})\leq E_{h}(\phi^{0}) for any m+m\in\mathbb{Z}^{+}. On the other hand, from the definition of EhE_{h}, (2.10) and the LhL^{\infty}_{h}-bound (3.1) for ϕm\phi^{m}, we can deduce that

Eh(ϕm)\displaystyle E_{h}(\phi^{m}) =ϵ42Δhϕm22+12β(ϕm)22+(λ22+λϵpη2)ϕm22(ϵ2+pη2+λϵ2)hϕm22\displaystyle=\frac{\epsilon^{4}}{2}\|\Delta_{h}\phi^{m}\|_{2}^{2}+\frac{1}{2}\|\beta(\phi^{m})\|_{2}^{2}+\left(\frac{\lambda^{2}}{2}+\frac{\lambda\epsilon^{p}\eta}{2}\right)\|\phi^{m}\|_{2}^{2}-\left(\frac{\epsilon^{2+p}\eta}{2}+\lambda\epsilon^{2}\right)\|\nabla_{h}\phi^{m}\|_{2}^{2}
λϕm,β(ϕm)Ω+ϵ2β(ϕm),ζ=x,y,zaζ(|Dζϕm|2)ΩϵpηB(ϕm),1Ω\displaystyle\quad-\lambda\left<\phi^{m},\beta(\phi^{m})\right>_{\Omega}+\epsilon^{2}\left<\beta^{\prime}(\phi^{m}),\sum_{\zeta=x,y,z}a_{\zeta}\left(\left|D_{\zeta}\phi^{m}\right|^{2}\right)\right>_{\Omega}-\epsilon^{p}\eta\left<B(\phi^{m}),1\right>_{\Omega}
ϵ42Δhϕm22+14β(ϕm)22C(1+hϕm22)\displaystyle\geq\frac{\epsilon^{4}}{2}\|\Delta_{h}\phi^{m}\|_{2}^{2}+\frac{1}{4}\|\beta(\phi^{m})\|_{2}^{2}-C(1+\|\nabla_{h}\phi^{m}\|_{2}^{2})
ϵ42Δhϕm22+14β(ϕm)22C(1+Δhϕm2ϕm2)\displaystyle\geq\frac{\epsilon^{4}}{2}\|\Delta_{h}\phi^{m}\|_{2}^{2}+\frac{1}{4}\|\beta(\phi^{m})\|_{2}^{2}-C(1+\|\Delta_{h}\phi^{m}\|_{2}\|\phi^{m}\|_{2})
ϵ44Δhϕm22+14β(ϕm)22C,\displaystyle\geq\frac{\epsilon^{4}}{4}\|\Delta_{h}\phi^{m}\|_{2}^{2}+\frac{1}{4}\|\beta(\phi^{m})\|_{2}^{2}-C,

where C>0C>0 may depend on λ\lambda, η\eta, ϵ\epsilon, |Ω||\Omega|, but is independent of mm. Thus, as long as the initial discrete energy Eh(ϕ0)E_{h}(\phi^{0}) is finite, it holds that Δhϕm2C\|\Delta_{h}\phi^{m}\|_{2}\leq C for any m+m\in\mathbb{Z}^{+}. This estimate combined with the fact ϕm<1\|\phi^{m}\|_{\infty}<1 leads to the conclusion (4.2). The proof is complete. ∎

5 Optimal Rate Convergence Analysis and Error Estimate

In this section, we perform the convergence analysis and derive error estimates for the discrete approximation of solutions to the FCH equation (1.1).

For the continuous problem (1.1) in the periodic setting with Ω=(0,1)d\Omega=(0,1)^{d}, 1d31\leq d\leq 3, existence and uniqueness of a global weak solution denoted by Φ\Phi has already been established in Proposition 2.1, provided that its initial datum Φ0\Phi_{0} satisfies Φ0Hper2(Ω)\Phi_{0}\in H^{2}_{\mathrm{per}}(\Omega), β(Φ0)Lper2(Ω)\beta(\Phi_{0})\in L^{2}_{\mathrm{per}}(\Omega) and ΩΦ0dx(1,1)\int_{\Omega}\Phi_{0}\mathrm{d}x\in(-1,1). Besides, in view of Proposition 2.2 and the discussions made in Remark 2.1, hereafter we assume that the initial datum Φ0\Phi_{0} is smooth enough and strictly separated from ±1\pm 1 so that the corresponding solution Φ\Phi to the continuous problem (1.1) belongs to the regularity class \mathcal{R} such that

Φ:=H4(0,T;Hper2(Ω))H2(0,T;Hper8(Ω))C([0,T];Hper12(Ω)),\Phi\in\mathcal{R}:=H^{4}(0,T;H^{2}_{\mathrm{per}}(\Omega))\cap H^{2}(0,T;H_{\mathrm{per}}^{8}(\Omega))\cap C([0,T];H_{\mathrm{per}}^{12}(\Omega)), (5.1)

on [0,T][0,T] for some T>0T>0. Moreover, we assume that the following strict separation property holds:

1+θ0Φ(x,t)1θ0,(x,t)Ω¯×[0,T],-1+\theta_{0}\leq\Phi(x,t)\leq 1-\theta_{0},\quad\forall\,(x,t)\in\overline{\Omega}\times[0,T], (5.2)

where θ0(0,1)\theta_{0}\in\left(0,1\right). The assumption (5.1) on the smoothness of the exact solution Φ\Phi may not be optimal, but they are sufficient for the subsequent convergence analysis.

Adapting the notations in [29, 55], we introduce the projection operator 𝒫N:Cper(Ω¯)K\mathcal{P}_{N}:C_{\mathrm{per}}(\overline{\Omega})\to\mathcal{B}^{K}, which is the space of trigonometric polynomials of degree less than or equal to K=[N/2]K=[N/2]. Define

ΦN(,t):=𝒫NΦ(,t)K\Phi_{N}(\cdot,t):=\mathcal{P}_{N}\Phi(\cdot,t)\in\mathcal{B}^{K}

the (spatial) Fourier projection of the exact solution Φ\Phi. If ΦL(0,T;Hperl(Ω))\Phi\in L^{\infty}(0,T;H_{\mathrm{per}}^{l}(\Omega)) for some l+l\in\mathbb{Z}^{+}, the following estimate is well-known (see [62]):

ΦNΦL(0,T;Hk(Ω))ChlkΦL(0,T;Hl(Ω)), 0kl.\|\Phi_{N}-\Phi\|_{L^{\infty}(0,T;H^{k}(\Omega))}\leq Ch^{l-k}\|\Phi\|_{L^{\infty}(0,T;H^{l}(\Omega))},\quad\forall\,0\leq k\leq l. (5.3)

In particular, from (5.1) and the Sobolev embedding Hper2(Ω)Cper(Ω¯)H_{\mathrm{per}}^{2}(\Omega)\hookrightarrow C_{\mathrm{per}}(\overline{\Omega}) in three dimensions, we can take k=2k=2, l=3l=3 in (5.3) and hh sufficiently small (i.e., NN is sufficiently large) so that the Fourier projection ΦN\Phi_{N} still preserves the strict separation property (however, with a relaxed bound):

1+θ02ΦN(x,t)1θ02,(x,t)Ω¯×[0,T].-1+\frac{\theta_{0}}{2}\leq\Phi_{N}(x,t)\leq 1-\frac{\theta_{0}}{2},\quad\forall\,(x,t)\in\overline{\Omega}\times[0,T]. (5.4)

Given an arbitrary positive integer MM, we set the uniform time step as Δt=T/M>0\Delta t=T/M>0. For every mm\in\mathbb{N} with mMm\leq M, we set

Φm=Φ(,tm)andΦNm=ΦN(,tm),withtm=mΔt.\Phi^{m}=\Phi(\cdot,t_{m})\quad\text{and}\quad\Phi_{N}^{m}=\Phi_{N}(\cdot,t_{m}),\quad\text{with}\ t_{m}=m\Delta t.

Since the exact solution Φ\Phi preserves the mass (recall Proposition 2.1) and ΦNK\Phi_{N}\in\mathcal{B}^{K}, it is straightforward to check that for all 0mM10\leq m\leq M-1, the property of mass conservation also holds for the Fourier projection ΦN\Phi_{N} at the discrete level of time:

ΩΦN(,tm+1)dx=ΩΦ(,tm+1)dx=ΩΦ(,tm)dx=ΩΦN(,tm)dx.\int_{\Omega}\Phi_{N}(\cdot,t_{m+1})\,\mathrm{d}x=\int_{\Omega}\Phi(\cdot,t_{m+1})\,\mathrm{d}x=\int_{\Omega}\Phi(\cdot,t_{m})\,\mathrm{d}x=\int_{\Omega}\Phi_{N}(\cdot,t_{m})\,\mathrm{d}x. (5.5)

Next, we define the canonical grid projection operator 𝒫h:Cper(Ω¯)𝒞per\mathcal{P}_{h}:C_{\mathrm{per}}(\overline{\Omega})\to\mathcal{C}_{\mathrm{per}}. For any gCper(Ω¯)g\in C_{\mathrm{per}}(\overline{\Omega}), we set g~=𝒫hg𝒞per\widetilde{g}=\mathcal{P}_{h}{g}\in\mathcal{C}_{\mathrm{per}} with grid values

g~i,j,k=g((i12)h,(j12)h,(k12)h),1i,j,kN.\widetilde{g}_{i,j,k}=g\Big{(}\big{(}i-\frac{1}{2}\big{)}h,\big{(}j-\frac{1}{2}\big{)}h,\big{(}k-\frac{1}{2}\big{)}h\Big{)},\quad 1\leq i,j,k\leq N.

Then for the fully discrete system (2.18)–(2.19), we take its initial datum as ϕ0=𝒫hΦN0=𝒫h(𝒫NΦ0)\phi^{0}=\mathcal{P}_{h}\Phi_{N}^{0}=\mathcal{P}_{h}(\mathcal{P}_{N}\Phi_{0}), i.e.,

ϕi,j,k0:=ΦN((i12)h,(j12)h,(k12)h,0),1i,j,kN.\phi_{i,j,k}^{0}:=\Phi_{N}\Big{(}\big{(}i-\frac{1}{2}\big{)}h,\big{(}j-\frac{1}{2}\big{)}h,\big{(}k-\frac{1}{2}\big{)}h,0\Big{)},\quad 1\leq i,j,k\leq N. (5.6)

The error grid function that we are going to investigate is given by

em:=𝒫hΦNmϕm,m,mM.e^{m}:=\mathcal{P}_{h}\Phi_{N}^{m}-\phi^{m},\qquad\forall\,m\in\mathbb{N},\ m\leq M. (5.7)

Keeping the definitions of 𝒫N\mathcal{P}_{N} and 𝒫h\mathcal{P}_{h} in mind, we infer from (5.5), (5.6) and Proposition 2.4 that

𝒫hΦNm¯=1|Ω|ΩΦNmdx=1|Ω|ΩΦN0dx=𝒫hΦN0¯=ϕ0¯=ϕm¯.\overline{\mathcal{P}_{h}\Phi_{N}^{m}}=\frac{1}{|\Omega|}\int_{\Omega}\Phi_{N}^{m}\,\mathrm{d}x=\frac{1}{|\Omega|}\int_{\Omega}\Phi_{N}^{0}\,\mathrm{d}x=\overline{\mathcal{P}_{h}\Phi_{N}^{0}}=\overline{\phi^{0}}=\overline{\phi^{m}}. (5.8)

As a consequence, we have

em¯=0,m,mM.\overline{e^{m}}=0,\quad\forall\,m\in\mathbb{N},\ m\leq M.

The main result of this section reads as follows

Theorem 5.3 (Error estimate).

Suppose that the exact solution Φ\Phi of the FCH equation (1.1) belongs to the regularity class \mathcal{R} (see (5.1)) and satisfies the strict separation property (5.2). Then, provided that Δt\Delta t and hh are sufficiently small, and under the linear refinement requirement ΔtC1h\Delta t\leq C_{1}h for some fixed C1>0C_{1}>0, we have

em2+[ϵ4Δtk=1m(hΔhek22+Δhek22)]12C(Δt+h2),\|e^{m}\|_{2}+\left[\epsilon^{4}\Delta t\sum_{k=1}^{m}\Big{(}\|\nabla_{h}\Delta_{h}e^{k}\|_{2}^{2}+\|\Delta_{h}e^{k}\|_{2}^{2}\Big{)}\right]^{\frac{1}{2}}\leq C(\Delta t+h^{2}), (5.9)

for any m+m\in\mathbb{Z}^{+} such that tm=mΔtTt_{m}=m\Delta t\leq T, where eme^{m} is defined as in (5.7) and the constant C>0C>0 is independent of mm, Δt\Delta t and hh.

Remark 5.1.

The linear refinement requirement ΔtC1h\Delta t\leq C_{1}h on the size of time step is a technical assumption due to the usage of inverse inequalities between norms (see (5.30) and (5.36) below).

5.1 Higher-order consistency analysis

First, applying the temporal discretization, we see that the Fourier projection ΦN\Phi_{N} solves the following semi-implicit equation:

ΦNn+1ΦNnΔt\displaystyle\frac{\Phi_{N}^{n+1}-\Phi_{N}^{n}}{\Delta t} =Δ[ϵ4Δ2ΦNn+1+β(ΦNn+1)β(ΦNn+1)+ϵ2β′′(ΦNn+1)|ΦNn+1|2\displaystyle=\Delta\Big{[}\epsilon^{4}\Delta^{2}\Phi_{N}^{n+1}+\beta(\Phi_{N}^{n+1})\beta^{\prime}(\Phi_{N}^{n+1})+\epsilon^{2}\beta^{\prime\prime}(\Phi_{N}^{n+1})|\nabla\Phi_{N}^{n+1}|^{2}
2ϵ2(β(ΦNn+1)ΦNn+1)+λ(λ+ϵpη)ΦNn+1λΦNnβ(ΦNn)\displaystyle\quad-2\epsilon^{2}\nabla\cdot\left(\beta^{\prime}(\Phi_{N}^{n+1})\nabla\Phi_{N}^{n+1}\right)+\lambda(\lambda+\epsilon^{p}\eta)\Phi_{N}^{n+1}-\lambda\Phi_{N}^{n}\beta^{\prime}(\Phi_{N}^{n})
+ϵ2(2λ+ϵpη)ΔΦNn(λ+ϵpη)β(ΦNn)]+F~Nn+1,\displaystyle\quad+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta\Phi_{N}^{n}-(\lambda+\epsilon^{p}\eta)\beta(\Phi_{N}^{n})\Big{]}+\widetilde{F}_{N}^{n+1}, (5.10)

where the truncation error is given by

F~Nn+1\displaystyle\widetilde{F}_{N}^{n+1} =(ΦNn+1ΦNnΔttΦ)+ϵ4Δ3(ΦΦNn+1)+Δ[β(Φ)β(Φ)β(ΦNn+1)β(ΦNn+1)]\displaystyle=\left(\frac{\Phi_{N}^{n+1}-\Phi_{N}^{n}}{\Delta t}-\partial_{t}\Phi\right)+\epsilon^{4}\Delta^{3}(\Phi-\Phi_{N}^{n+1})+\Delta\big{[}\beta(\Phi)\beta^{\prime}(\Phi)-\beta(\Phi_{N}^{n+1})\beta^{\prime}(\Phi_{N}^{n+1})\big{]}
+ϵ2Δ[β′′(Φ)|Φ|2β′′(ΦNn+1)|ΦNn+1|2]2ϵ2Δ2[β(Φ)β(ΦNn+1)]\displaystyle\quad+\epsilon^{2}\Delta\big{[}\beta^{\prime\prime}(\Phi)|\nabla\Phi|^{2}-\beta^{\prime\prime}(\Phi_{N}^{n+1})|\nabla\Phi_{N}^{n+1}|^{2}\big{]}-2\epsilon^{2}\Delta^{2}\big{[}\beta(\Phi)-\beta(\Phi_{N}^{n+1})\big{]}
+λ(λ+ϵpη)Δ(ΦΦNn+1)λΔ[Φβ(Φ)ΦNnβ(ΦNn)]\displaystyle\quad+\lambda(\lambda+\epsilon^{p}\eta)\Delta(\Phi-\Phi_{N}^{n+1})-\lambda\Delta\big{[}\Phi\beta^{\prime}(\Phi)-\Phi_{N}^{n}\beta^{\prime}(\Phi_{N}^{n})\big{]}
+ϵ2(2λ+ϵpη)Δ2(ΦΦNn)(λ+ϵpη)Δ[β(Φ)β(ΦNn)].\displaystyle\quad+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta^{2}(\Phi-\Phi_{N}^{n})-(\lambda+\epsilon^{p}\eta)\Delta\big{[}\beta(\Phi)-\beta(\Phi_{N}^{n})\big{]}.

Recalling the assumptions (5.1)–(5.2), then using Taylor’s expansion in time and the error estimate (5.3) (e.g., with l=11l=11, k=8k=8), we have

F~Nn+1=ΔtFN,1n+1+O(Δt2)+O(h3),\widetilde{F}_{N}^{n+1}=\Delta tF_{N,1}^{n+1}+O(\Delta t^{2})+O(h^{3}),

where the spatial function FN,1n+1F_{N,1}^{n+1} only depends on ΦN\Phi_{N}, Φ\Phi, and it is sufficiently smooth in the sense that its derivatives are bounded. From the mass conservation (5.5) and the periodic boundary condition, we infer that ΩF~Nn+1dx=0\int_{\Omega}\widetilde{F}_{N}^{n+1}\,\mathrm{d}x=0. This fact combined with Taylor’s expansion further implies ΩFN,1n+1dx=0\int_{\Omega}F_{N,1}^{n+1}\,\mathrm{d}x=0.

Performing a further spatial discretization to (5.10) and applying a careful consistency analysis via Taylor’s expansion, we can actually show that 𝒫hΦNn+1\mathcal{P}_{h}\Phi_{N}^{n+1} solves the discrete system (2.18)–(2.19) with a first order accuracy in time and a second order accuracy in space. However, a restrictive refinement condition like Ch2ΔtC′′h2C^{\prime}h^{2}\leq\Delta t\leq C^{\prime\prime}h^{2} for some C′′>C>0C^{\prime\prime}>C^{\prime}>0 is needed to complete the proof. The detailed calculations are left to the interested readers.

In order to prove Theorem 5.3 under a weaker (i.e., linear) refinement condition on the time step size Δt\Delta t, we perform a higher-order consistency analysis for the auxiliary profile

Φ^=ΦN+Δt𝒫NΦΔt,1+Δt2𝒫NΦΔt,2+h2𝒫NΦh,1.\widehat{\Phi}=\Phi_{N}+\Delta t\mathcal{P}_{N}\Phi_{\Delta t,1}+\Delta t^{2}\mathcal{P}_{N}\Phi_{\Delta t,2}+h^{2}\mathcal{P}_{N}\Phi_{h,1}. (5.11)

Here we note that the centered difference used in the spatial discretization gives local truncation errors with only even order terms. In this manner, we find that a higher order consistency of O(Δt3+h3)O(\Delta t^{3}+h^{3}) holds for the numerical scheme (2.18)–(2.19) with 𝒫hΦ^n+1\mathcal{P}_{h}\widehat{\Phi}^{n+1}, see (5.23)–(5.24) below. The approximate expansion (5.11) enables us to derive a higher order convergence estimate of the modified numerical error function e^n\widehat{e}^{n} (see (5.28)) between the constructed profile 𝒫hΦ^n\mathcal{P}_{h}\widehat{\Phi}^{n} and the numerical solution ϕn\phi^{n} in the discrete 1,h\|\cdot\|_{-1,h} norm (see (5.35)). This combined with the inverse inequality eventually yields a refined \|\cdot\|_{\infty} bound of the numerical solution ϕn\phi^{n}, more precisely, the property of strict separation from pure states ±1\pm 1 (see (5.31)) that is uniform with respect to the time step and the grid size (cf. Remark 3.1).

The supplementary field ΦΔt,1\Phi_{\Delta t,1} in the profile (5.11) only depends on the exact solution Φ\Phi and can be constructed via a perturbation expansion argument (see e.g., [55, 63]). Indeed, we find the temporal correction ΦΔt,1\Phi_{\Delta t,1} by solving the following linear equation:

tΦΔt,1\displaystyle\partial_{t}\Phi_{\Delta t,1} =Δ[ϵ4Δ2ΦΔt,1+(β(ΦN)2+β(ΦN)β′′(ΦN))ΦΔt,1\displaystyle=\Delta\bigg{[}\epsilon^{4}\Delta^{2}\Phi_{\Delta t,1}+\left(\beta^{\prime}(\Phi_{N})^{2}+\beta(\Phi_{N})\beta^{\prime\prime}(\Phi_{N})\right)\Phi_{\Delta t,1}
+ϵ2(2β′′(ΦN)ΦNΦΔt,1+β′′′(ΦN)|ΦN|2ΦΔt,1)\displaystyle\qquad+\epsilon^{2}\left(2\beta^{\prime\prime}(\Phi_{N})\nabla\Phi_{N}\cdot\nabla\Phi_{\Delta t,1}+\beta^{\prime\prime\prime}(\Phi_{N})|\nabla\Phi_{N}|^{2}\Phi_{\Delta t,1}\right)
2ϵ2(β(ΦN)ΦΔt,1+β′′(ΦN)ΦΔt,1ΦN)\displaystyle\qquad-2\epsilon^{2}\nabla\cdot\left(\beta^{\prime}(\Phi_{N})\nabla\Phi_{\Delta t,1}+\beta^{\prime\prime}(\Phi_{N})\Phi_{\Delta t,1}\nabla\Phi_{N}\right)
+λ(λ+ϵpη)ΦΔt,1λ(ΦNβ′′(ΦN)+β(ΦN))ΦΔt,1\displaystyle\qquad+\lambda(\lambda+\epsilon^{p}\eta)\Phi_{\Delta t,1}-\lambda\left(\Phi_{N}\beta^{\prime\prime}(\Phi_{N})+\beta^{\prime}(\Phi_{N})\right)\Phi_{\Delta t,1}
+ϵ2(2λ+ϵpη)ΔΦΔt,1(λ+ϵpη)β(ΦN)ΦΔt,1]FN,1,\displaystyle\qquad+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta\Phi_{\Delta t,1}-(\lambda+\epsilon^{p}\eta)\beta^{\prime}(\Phi_{N})\Phi_{\Delta t,1}\bigg{]}-F_{N,1}, (5.12)

subject to the zero initial condition ΦΔt,1(,0)=0\Phi_{\Delta t,1}(\cdot,0)=0 and periodic boundary conditions. Once the Fourier projection ΦN\Phi_{N} is given, we take the external source term FN,1F_{N,1} in (5.12) to be a smooth function satisfying

FN,1(,tn)=FN,1nandΩFN,1(,t)dx=0,t[0,T].F_{N,1}(\cdot,t_{n})=F_{N,1}^{n}\quad\text{and}\quad\int_{\Omega}F_{N,1}(\cdot,t)\,\mathrm{d}x=0,\quad\forall\,t\in[0,T].

The existence and uniqueness of a solution to the linear sixth order parabolic equation (5.12) on [0,T][0,T] can be easily proved, for instance, by using the Faedo-Galerkin method. In particular, the solution ΦΔt,1\Phi_{\Delta t,1} only depends on ΦN\Phi_{N}, and its derivatives in various orders are bounded (recalling that ΦN\Phi_{N} is sufficiently smooth). Moreover, it satisfies the zero mass property:

ΩΦΔt,1(,t)dx=ΩΦΔt,1(,0)dx=0,t[0,T].\int_{\Omega}\Phi_{\Delta t,1}(\cdot,t)\,\mathrm{d}x=\int_{\Omega}\Phi_{\Delta t,1}(\cdot,0)\,\mathrm{d}x=0,\quad\forall\,t\in[0,T].

Next, an application of the semi-implicit discretization for (5.12) yields

ΦΔt,1n+1ΦΔt,1nΔt\displaystyle\frac{\Phi_{\Delta t,1}^{n+1}-\Phi_{\Delta t,1}^{n}}{\Delta t} =Δ[ϵ4Δ2ΦΔt,1n+1+(β(ΦNn+1)2+β(ΦNn+1)β′′(ΦNn+1))ΦΔt,1n+1\displaystyle=\Delta\bigg{[}\epsilon^{4}\Delta^{2}\Phi_{\Delta t,1}^{n+1}+\left(\beta^{\prime}(\Phi_{N}^{n+1})^{2}+\beta(\Phi_{N}^{n+1})\beta^{\prime\prime}(\Phi_{N}^{n+1})\right)\Phi_{\Delta t,1}^{n+1}
+ϵ2(2β′′(ΦNn+1)ΦNn+1ΦΔt,1n+1+β′′′(ΦNn+1)|ΦNn+1|2ΦΔt,1n+1)\displaystyle\qquad+\epsilon^{2}\left(2\beta^{\prime\prime}(\Phi_{N}^{n+1})\nabla\Phi_{N}^{n+1}\cdot\nabla\Phi_{\Delta t,1}^{n+1}+\beta^{\prime\prime\prime}(\Phi_{N}^{n+1})|\nabla\Phi_{N}^{n+1}|^{2}\Phi_{\Delta t,1}^{n+1}\right)
2ϵ2(β(ΦNn+1)ΦΔt,1n+1+β′′(ΦNn+1)ΦΔt,1n+1ΦNn+1)\displaystyle\qquad-2\epsilon^{2}\nabla\cdot\left(\beta^{\prime}(\Phi_{N}^{n+1})\nabla\Phi_{\Delta t,1}^{n+1}+\beta^{\prime\prime}(\Phi_{N}^{n+1})\Phi_{\Delta t,1}^{n+1}\nabla\Phi_{N}^{n+1}\right)
+λ(λ+ϵpη)ΦΔt,1n+1λ(ΦNnβ′′(ΦNn)+β(ΦNn))ΦΔt,1n\displaystyle\qquad+\lambda(\lambda+\epsilon^{p}\eta)\Phi_{\Delta t,1}^{n+1}-\lambda\left(\Phi_{N}^{n}\beta^{\prime\prime}(\Phi_{N}^{n})+\beta^{\prime}(\Phi_{N}^{n})\right)\Phi_{\Delta t,1}^{n}
+ϵ2(2λ+ϵpη)ΔΦΔt,1n(λ+ϵpη)β(ΦNn)ΦΔt,1n]FN,1n+1+O(Δt),\displaystyle\qquad+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta\Phi_{\Delta t,1}^{n}-(\lambda+\epsilon^{p}\eta)\beta^{\prime}(\Phi_{N}^{n})\Phi_{\Delta t,1}^{n}\bigg{]}-F_{N,1}^{n+1}+O(\Delta t), (5.13)

where we denote by ΦΔt,1m=ΦΔt,1(,tm)\Phi_{\Delta t,1}^{m}=\Phi_{\Delta t,1}(\cdot,t_{m}), tm=mΔtt_{m}=m\Delta t, mm\in\mathbb{N}, mMm\leq M. At the discrete level of time, we note that for every nn\in\mathbb{N}, it also holds ΩΦΔt,1n+1dx=ΩΦΔt,1ndx\int_{\Omega}\Phi_{\Delta t,1}^{n+1}\,\mathrm{d}x=\int_{\Omega}\Phi_{\Delta t,1}^{n}\,\mathrm{d}x. This fact further implies

Ω𝒫NΦΔt,1mdx=0,m,mM.\int_{\Omega}\mathcal{P}_{N}\Phi_{\Delta t,1}^{m}\,\mathrm{d}x=0,\quad\forall\,m\in\mathbb{N},\ m\leq M. (5.14)

Combining (5.10) and (5.13), we can obtain the following second order in temporal space truncation error for

Φ^1:=ΦN+Δt𝒫NΦΔt,1\widehat{\Phi}_{1}:=\Phi_{N}+\Delta t\mathcal{P}_{N}\Phi_{\Delta t,1}

such that

Φ^1n+1Φ^1nΔt=Δ\displaystyle\frac{\widehat{\Phi}_{1}^{n+1}-\widehat{\Phi}_{1}^{n}}{\Delta t}=\Delta [ϵ4Δ2Φ^1n+1+β(Φ^1n+1)β(Φ^1n+1)+ϵ2β′′(Φ^1n+1)|Φ^1n+1|2\displaystyle\Big{[}\epsilon^{4}\Delta^{2}\widehat{\Phi}_{1}^{n+1}+\beta(\widehat{\Phi}_{1}^{n+1})\beta^{\prime}(\widehat{\Phi}_{1}^{n+1})+\epsilon^{2}\beta^{\prime\prime}(\widehat{\Phi}_{1}^{n+1})|\nabla\widehat{\Phi}_{1}^{n+1}|^{2}
2ϵ2(β(Φ^1n+1)Φ^1n+1)+λ(λ+ϵpη)Φ^n+1λΦ^1nβ(Φ^1n)\displaystyle-2\epsilon^{2}\nabla\cdot\left(\beta^{\prime}(\widehat{\Phi}_{1}^{n+1})\nabla\widehat{\Phi}_{1}^{n+1}\right)+\lambda(\lambda+\epsilon^{p}\eta)\widehat{\Phi}^{n+1}-\lambda\widehat{\Phi}_{1}^{n}\beta^{\prime}(\widehat{\Phi}_{1}^{n})
+ϵ2(2λ+ϵpη)ΔΦ^1n(λ+ϵpη)β(Φ^n)]+Δt2FN,2n+1+O(Δt3)+O(h3).\displaystyle+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta\widehat{\Phi}_{1}^{n}-(\lambda+\epsilon^{p}\eta)\beta(\widehat{\Phi}^{n})\Big{]}+\Delta t^{2}F_{N,2}^{n+1}+O(\Delta t^{3})+O(h^{3}). (5.15)

Following arguments similar to those in [55, 64, 65], we can further derive the second order temporal correction ΦΔt,2\Phi_{\Delta t,2} as well as the spatial correction Φh,1\Phi_{h,1}, respectively.

To this end, the next order temporal correction function ΦΔt,2\Phi_{\Delta t,2} is given by the linear equation

tΦΔt,2\displaystyle\partial_{t}\Phi_{\Delta t,2} =Δ[ϵ4Δ2ΦΔt,2+(β(Φ^1)2+β(Φ^1)β′′(Φ^1))ΦΔt,2\displaystyle=\Delta\bigg{[}\epsilon^{4}\Delta^{2}\Phi_{\Delta t,2}+\left(\beta^{\prime}(\widehat{\Phi}_{1})^{2}+\beta(\widehat{\Phi}_{1})\beta^{\prime\prime}(\widehat{\Phi}_{1})\right)\Phi_{\Delta t,2}
+ϵ2(2β′′(Φ^1)Φ^1ΦΔt,2+β′′′(Φ^1)|Φ^1|2ΦΔt,2)\displaystyle\qquad+\epsilon^{2}\left(2\beta^{\prime\prime}(\widehat{\Phi}_{1})\nabla\widehat{\Phi}_{1}\cdot\nabla\Phi_{\Delta t,2}+\beta^{\prime\prime\prime}(\widehat{\Phi}_{1})|\nabla\widehat{\Phi}_{1}|^{2}\Phi_{\Delta t,2}\right)
2ϵ2(β(Φ^1)ΦΔt,2+β′′(Φ^1)ΦΔt,2Φ^1)\displaystyle\qquad-2\epsilon^{2}\nabla\cdot\left(\beta^{\prime}(\widehat{\Phi}_{1})\nabla\Phi_{\Delta t,2}+\beta^{\prime\prime}(\widehat{\Phi}_{1})\Phi_{\Delta t,2}\nabla\widehat{\Phi}_{1}\right)
+λ(λ+ϵpη)ΦΔt,2λ(Φ^1β′′(Φ^1)+β(Φ^1))ΦΔt,2\displaystyle\qquad+\lambda(\lambda+\epsilon^{p}\eta)\Phi_{\Delta t,2}-\lambda\left(\widehat{\Phi}_{1}\beta^{\prime\prime}(\widehat{\Phi}_{1})+\beta^{\prime}(\widehat{\Phi}_{1})\right)\Phi_{\Delta t,2}
+ϵ2(2λ+ϵpη)ΔΦΔt,2(λ+ϵpη)β(Φ^1)ΦΔt,2]FN,2,\displaystyle\qquad+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta\Phi_{\Delta t,2}-(\lambda+\epsilon^{p}\eta)\beta^{\prime}(\widehat{\Phi}_{1})\Phi_{\Delta t,2}\bigg{]}-F_{N,2}, (5.16)

subject to the zero initial condition ΦΔt,2(,0)=0\Phi_{\Delta t,2}(\cdot,0)=0 and periodic boundary conditions. Like before, the solution ΦΔt,2\Phi_{\Delta t,2} depends only on the exact solution Φ\Phi and its derivatives of various orders are bounded. An application of the semi-implicit discretization to (5.16) implies that

ΦΔt,2n+1ΦΔt,2nΔt\displaystyle\frac{\Phi_{\Delta t,2}^{n+1}-\Phi_{\Delta t,2}^{n}}{\Delta t} =Δ[ϵ4Δ2ΦΔt,2n+1+(β(Φ^1n+1)2+β(Φ^1n+1)β′′(Φ^1n+1))ΦΔt,2n+1\displaystyle=\Delta\bigg{[}\epsilon^{4}\Delta^{2}\Phi_{\Delta t,2}^{n+1}+\left(\beta^{\prime}(\widehat{\Phi}_{1}^{n+1})^{2}+\beta(\widehat{\Phi}_{1}^{n+1})\beta^{\prime\prime}(\widehat{\Phi}_{1}^{n+1})\right)\Phi_{\Delta t,2}^{n+1}
+ϵ2(2β′′(Φ^1n+1)Φ^1n+1ΦΔt,2n+1+β′′′(Φ^1n+1)|Φ^1n+1|2ΦΔt,2n+1)\displaystyle\qquad+\epsilon^{2}\left(2\beta^{\prime\prime}(\widehat{\Phi}_{1}^{n+1})\nabla\widehat{\Phi}_{1}^{n+1}\cdot\nabla\Phi_{\Delta t,2}^{n+1}+\beta^{\prime\prime\prime}(\widehat{\Phi}_{1}^{n+1})|\nabla\widehat{\Phi}_{1}^{n+1}|^{2}\Phi_{\Delta t,2}^{n+1}\right)
2ϵ2(β(Φ^1n+1)ΦΔt,2n+1+β′′(Φ^1n+1)ΦΔt,2n+1Φ^1n+1)\displaystyle\qquad-2\epsilon^{2}\nabla\cdot\left(\beta^{\prime}(\widehat{\Phi}_{1}^{n+1})\nabla\Phi_{\Delta t,2}^{n+1}+\beta^{\prime\prime}(\widehat{\Phi}_{1}^{n+1})\Phi_{\Delta t,2}^{n+1}\nabla\widehat{\Phi}_{1}^{n+1}\right)
+λ(λ+ϵpη)ΦΔt,2n+1λ(Φ^1nβ′′(Φ^1n)+β(Φ^1n))ΦΔt,2n\displaystyle\qquad+\lambda(\lambda+\epsilon^{p}\eta)\Phi_{\Delta t,2}^{n+1}-\lambda\left(\widehat{\Phi}_{1}^{n}\beta^{\prime\prime}(\widehat{\Phi}_{1}^{n})+\beta^{\prime}(\widehat{\Phi}_{1}^{n})\right)\Phi_{\Delta t,2}^{n}
+ϵ2(2λ+ϵpη)ΔΦΔt,2n(λ+ϵpη)β(Φ^1n)ΦΔt,2n]FN,2n+1+O(Δt).\displaystyle\qquad+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta\Phi_{\Delta t,2}^{n}-(\lambda+\epsilon^{p}\eta)\beta^{\prime}(\widehat{\Phi}_{1}^{n})\Phi_{\Delta t,2}^{n}\bigg{]}-F_{N,2}^{n+1}+O(\Delta t). (5.17)

Then a combination of (5.16) and (5.17) yields the third order temporal truncation error for the profile

Φ^2:=ΦN+Δt𝒫NΦΔt,1+Δt2𝒫NΦΔt,2\widehat{\Phi}_{2}:=\Phi_{N}+\Delta t\mathcal{P}_{N}\Phi_{\Delta t,1}+\Delta t^{2}\mathcal{P}_{N}\Phi_{\Delta t,2}

such that

Φ^2n+1Φ^2nΔt=Δ\displaystyle\frac{\widehat{\Phi}_{2}^{n+1}-\widehat{\Phi}_{2}^{n}}{\Delta t}=\Delta [ϵ4Δ2Φ^2n+1+β(Φ^2n+1)β(Φ^2n+1)+ϵ2β′′(Φ^2n+1)|Φ^2n+1|2\displaystyle\Big{[}\epsilon^{4}\Delta^{2}\widehat{\Phi}_{2}^{n+1}+\beta(\widehat{\Phi}_{2}^{n+1})\beta^{\prime}(\widehat{\Phi}_{2}^{n+1})+\epsilon^{2}\beta^{\prime\prime}(\widehat{\Phi}_{2}^{n+1})|\nabla\widehat{\Phi}_{2}^{n+1}|^{2}
2ϵ2(β(Φ^2n+1)Φ^2n+1)+λ(λ+ϵpη)Φ^2n+1λΦ^2nβ(Φ^2n)\displaystyle-2\epsilon^{2}\nabla\cdot\left(\beta^{\prime}(\widehat{\Phi}_{2}^{n+1})\nabla\widehat{\Phi}_{2}^{n+1}\right)+\lambda(\lambda+\epsilon^{p}\eta)\widehat{\Phi}_{2}^{n+1}-\lambda\widehat{\Phi}_{2}^{n}\beta^{\prime}(\widehat{\Phi}_{2}^{n})
+ϵ2(2λ+ϵpη)ΔΦ^2n(λ+ϵpη)β(Φ^2n)]+O(Δt3)+O(h3).\displaystyle+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta\widehat{\Phi}_{2}^{n}-(\lambda+\epsilon^{p}\eta)\beta(\widehat{\Phi}_{2}^{n})\Big{]}+O(\Delta t^{3})+O(h^{3}). (5.18)

Finally, we construct the spatial correction term Φh,1\Phi_{h,1} to upgrade the spatial accuracy order. The following truncation error analysis for the spatial discretization can be obtained by using a straightforward Taylor expansion for Φ^2\widehat{\Phi}_{2}:

Φ^2n+1Φ^2nΔt=Δh\displaystyle\frac{\widehat{\Phi}_{2}^{n+1}-\widehat{\Phi}_{2}^{n}}{\Delta t}=\Delta_{h} [ϵ4Δh2Φ^2n+1+β(Φ^2n+1)β(Φ^2n+1)+ϵ2ζ=x,y,zβ′′(Φ^2n+1)aζ(|DζΦ^2n+1|2)\displaystyle\Big{[}\epsilon^{4}\Delta_{h}^{2}\widehat{\Phi}_{2}^{n+1}+\beta(\widehat{\Phi}_{2}^{n+1})\beta^{\prime}(\widehat{\Phi}_{2}^{n+1})+\epsilon^{2}\sum_{\zeta=x,y,z}\beta^{\prime\prime}(\widehat{\Phi}_{2}^{n+1})a_{\zeta}\left(|D_{\zeta}\widehat{\Phi}_{2}^{n+1}|^{2}\right)
2ϵ2ζ=x,y,zdλ(Aζ(β(Φ^2n+1))DζΦ^2n+1)+λ(λ+ϵpη)Φ^2n+1\displaystyle-2\epsilon^{2}\sum_{\zeta=x,y,z}d_{\lambda}\left(A_{\zeta}\big{(}\beta^{\prime}(\widehat{\Phi}_{2}^{n+1})\big{)}D_{\zeta}\widehat{\Phi}_{2}^{n+1}\right)+\lambda(\lambda+\epsilon^{p}\eta)\widehat{\Phi}_{2}^{n+1}
λΦ^2nβ(Φ^2n)+ϵ2(2λ+ϵpη)ΔhΦ^2n(λ+ϵpη)β(Φ^2n)]\displaystyle-\lambda\widehat{\Phi}_{2}^{n}\beta^{\prime}(\widehat{\Phi}_{2}^{n})+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta_{h}\widehat{\Phi}_{2}^{n}-(\lambda+\epsilon^{p}\eta)\beta(\widehat{\Phi}_{2}^{n})\Big{]}
+h2HN,1n+1+O(Δt3)+O(h3).\displaystyle+h^{2}H_{N,1}^{n+1}+O(\Delta t^{3})+O(h^{3}). (5.19)

Subsequently, the spatial correction function Φh,1\Phi_{h,1} is given by solving the linear equation

tΦh,1\displaystyle\partial_{t}\Phi_{h,1} =Δ[ϵ4Δ2Φh,1+(β(Φ^2)2+β(Φ^2)β′′(Φ^2))Φh,1\displaystyle=\Delta\bigg{[}\epsilon^{4}\Delta^{2}\Phi_{h,1}+\left(\beta^{\prime}(\widehat{\Phi}_{2})^{2}+\beta(\widehat{\Phi}_{2})\beta^{\prime\prime}(\widehat{\Phi}_{2})\right)\Phi_{h,1}
+ϵ2(2β′′(Φ^2)Φ^2Φh,1+β′′′(Φ^2)|Φ^2|2Φh,1)\displaystyle\qquad+\epsilon^{2}\left(2\beta^{\prime\prime}(\widehat{\Phi}_{2})\nabla\widehat{\Phi}_{2}\cdot\nabla\Phi_{h,1}+\beta^{\prime\prime\prime}(\widehat{\Phi}_{2})|\nabla\widehat{\Phi}_{2}|^{2}\Phi_{h,1}\right)
2ϵ2(β(Φ^2)Φh,1+β′′(Φ^2)Φh,1Φ^2)\displaystyle\qquad-2\epsilon^{2}\nabla\cdot\left(\beta^{\prime}(\widehat{\Phi}_{2})\nabla\Phi_{h,1}+\beta^{\prime\prime}(\widehat{\Phi}_{2})\Phi_{h,1}\nabla\widehat{\Phi}_{2}\right)
+λ(λ+ϵpη)Φh,1λ(Φ^2β′′(Φ^2)+β(Φ^2))Φh,1\displaystyle\qquad+\lambda(\lambda+\epsilon^{p}\eta)\Phi_{h,1}-\lambda\left(\widehat{\Phi}_{2}\beta^{\prime\prime}(\widehat{\Phi}_{2})+\beta^{\prime}(\widehat{\Phi}_{2})\right)\Phi_{h,1}
+ϵ2(2λ+ϵpη)ΔΦh,1(λ+ϵpη)β(Φ^2)Φh,1]HN,1,\displaystyle\qquad+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta\Phi_{h,1}-(\lambda+\epsilon^{p}\eta)\beta^{\prime}(\widehat{\Phi}_{2})\Phi_{h,1}\bigg{]}-H_{N,1}, (5.20)

subject to the zero initial condition Φh,1(,0)=0\Phi_{h,1}(\cdot,0)=0 and periodic boundary conditions. Again, the solution Φh,1\Phi_{h,1} depends only on the (sufficiently) exact solution Φ\Phi, with bounded differences of various orders.

Combining the above arguments, we have thus constructed the required auxiliary profile Φ^\widehat{\Phi} (recall (5.11)).

Remark 5.2.

Since the temporal/spatial correction functions ΦΔt,1\Phi_{\Delta t,1}, ΦΔt,2\Phi_{\Delta t,2}, Φh,1\Phi_{h,1} are sufficiently smooth and bounded, applying the projection estimate (5.3) and the separation property (5.4) for ΦN\Phi_{N}, we can obtain a strict separation property for Φ^\widehat{\Phi} such that

1+θΦ^(x,t)1θ,(x,t)Ω¯×[0,T],-1+\theta\leq\widehat{\Phi}(x,t)\leq 1-\theta,\quad\forall\,(x,t)\in\overline{\Omega}\times[0,T], (5.21)

with θ=θ0/4\theta=\theta_{0}/4, provided that the time step Δt\Delta t and the grid size hh are sufficiently small. Such a uniform separation property will be useful in the subsequent convergence analysis.

An application of a full discretization to (5.20) yields that

Φh,1n+1Φh,1nΔt\displaystyle\frac{\Phi_{h,1}^{n+1}-\Phi_{h,1}^{n}}{\Delta t} =Δh[ϵ4Δh2Φh,1n+1+(β(Φ^2n+1)2+β(Φ^2n+1)β′′(Φ^2n+1))Φh,1n+1\displaystyle=\Delta_{h}\bigg{[}\epsilon^{4}\Delta_{h}^{2}\Phi_{h,1}^{n+1}+\left(\beta^{\prime}(\widehat{\Phi}_{2}^{n+1})^{2}+\beta(\widehat{\Phi}_{2}^{n+1})\beta^{\prime\prime}(\widehat{\Phi}_{2}^{n+1})\right)\Phi_{h,1}^{n+1}
+ϵ2(2β′′(Φ^2n+1)hΦ^2n+1hΦh,1n+1+β′′′(Φ^2n+1)|hΦ^2n+1|2Φh,1n+1)\displaystyle\qquad+\epsilon^{2}\left(2\beta^{\prime\prime}(\widehat{\Phi}_{2}^{n+1})\nabla_{h}\widehat{\Phi}_{2}^{n+1}\cdot\nabla_{h}\Phi_{h,1}^{n+1}+\beta^{\prime\prime\prime}(\widehat{\Phi}_{2}^{n+1})|\nabla_{h}\widehat{\Phi}_{2}^{n+1}|^{2}\Phi_{h,1}^{n+1}\right)
2ϵ2h(β(Φ^2n+1)hΦh,1n+1+β′′(Φ^2n+1)Φh,1n+1hΦ^2n+1)\displaystyle\qquad-2\epsilon^{2}\nabla_{h}\cdot\left(\beta^{\prime}(\widehat{\Phi}_{2}^{n+1})\nabla_{h}\Phi_{h,1}^{n+1}+\beta^{\prime\prime}(\widehat{\Phi}_{2}^{n+1})\Phi_{h,1}^{n+1}\nabla_{h}\widehat{\Phi}_{2}^{n+1}\right)
+λ(λ+ϵpη)Φh,1n+1λ(Φ^2nβ′′(Φ^2n)+β(Φ^2n))Φh,1n\displaystyle\qquad+\lambda(\lambda+\epsilon^{p}\eta)\Phi_{h,1}^{n+1}-\lambda\left(\widehat{\Phi}_{2}^{n}\beta^{\prime\prime}(\widehat{\Phi}_{2}^{n})+\beta^{\prime}(\widehat{\Phi}_{2}^{n})\right)\Phi_{h,1}^{n}
+ϵ2(2λ+ϵpη)ΔhΦh,1n(λ+ϵpη)β(Φ^2n)Φh,1n]HN,1n+1+O(Δt+h2).\displaystyle\qquad+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta_{h}\Phi_{h,1}^{n}-(\lambda+\epsilon^{p}\eta)\beta^{\prime}(\widehat{\Phi}_{2}^{n})\Phi_{h,1}^{n}\bigg{]}-H_{N,1}^{n+1}+O(\Delta t+h^{2}). (5.22)

Hence, for the profile Φ^=ΦN+Δt𝒫NΦΔt,1+Δt2𝒫NΦΔt,2+h2𝒫NΦh,1\widehat{\Phi}=\Phi_{N}+\Delta t\mathcal{P}_{N}\Phi_{\Delta t,1}+\Delta t^{2}\mathcal{P}_{N}\Phi_{\Delta t,2}+h^{2}\mathcal{P}_{N}\Phi_{h,1}, by a careful consistency analysis via Taylor’s expansion, we can eventually arrive at the following fully discrete scheme for its discrete counterpart 𝒫hΦ^n+1\mathcal{P}_{h}\widehat{\Phi}^{n+1} such that (for the sake of simplicity, below we simply drop the projection operator 𝒫h\mathcal{P}_{h} in 𝒫hΦ^n\mathcal{P}_{h}\widehat{\Phi}^{n} and 𝒫hΦ^n+1\mathcal{P}_{h}\widehat{\Phi}^{n+1})

Φ^n+1Φ^nΔt=Δh\displaystyle\frac{\widehat{\Phi}^{n+1}-\widehat{\Phi}^{n}}{\Delta t}=\Delta_{h} [ϵ4Δh2Φ^n+1+β(Φ^n+1)β(Φ^n+1)+ϵ2ζ=x,y,zβ′′(Φ^n+1)aζ(|DζΦ^n+1|2)\displaystyle\Big{[}\epsilon^{4}\Delta_{h}^{2}\widehat{\Phi}^{n+1}+\beta(\widehat{\Phi}^{n+1})\beta^{\prime}(\widehat{\Phi}^{n+1})+\epsilon^{2}\sum_{\zeta=x,y,z}\beta^{\prime\prime}(\widehat{\Phi}^{n+1})a_{\zeta}\left(|D_{\zeta}\widehat{\Phi}^{n+1}|^{2}\right)
2ϵ2ζ=x,y,zdλ(Aζ(β(Φ^n+1))DζΦ^n+1)+λ(λ+ϵpη)Φ^n+1\displaystyle-2\epsilon^{2}\sum_{\zeta=x,y,z}d_{\lambda}\left(A_{\zeta}\big{(}\beta^{\prime}(\widehat{\Phi}^{n+1})\big{)}D_{\zeta}\widehat{\Phi}^{n+1}\right)+\lambda(\lambda+\epsilon^{p}\eta)\widehat{\Phi}^{n+1}
λΦ^nβ(Φ^n)+ϵ2(2λ+ϵpη)ΔhΦ^n(λ+ϵpη)β(Φ^n)]+τ^n+1,\displaystyle-\lambda\widehat{\Phi}^{n}\beta^{\prime}(\widehat{\Phi}^{n})+\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\Delta_{h}\widehat{\Phi}^{n}-(\lambda+\epsilon^{p}\eta)\beta(\widehat{\Phi}^{n})\Big{]}+\widehat{\tau}^{n+1}, (5.23)

with the higher order truncation error

τ^n+12C(Δt3+h3).\|\widehat{\tau}^{n+1}\|_{2}\leq C(\Delta t^{3}+h^{3}). (5.24)
Remark 5.3.

We note that the profile Φ^=ΦN+Δt𝒫NΦΔt,1+Δt2𝒫NΦΔt,2+h2𝒫NΦh,1K\widehat{\Phi}=\Phi_{N}+\Delta t\mathcal{P}_{N}\Phi_{\Delta t,1}+\Delta t^{2}\mathcal{P}_{N}\Phi_{\Delta t,2}+h^{2}\mathcal{P}_{N}\Phi_{h,1}\in\mathcal{B}^{K} still satisfies the mass conservation property at the discrete level of both time and space. Indeed, from its definition, we observe that 𝒫hΦ^0=𝒫hΦN0=ϕ0\mathcal{P}_{h}\widehat{\Phi}^{0}=\mathcal{P}_{h}\Phi_{N}^{0}=\phi^{0}. Then from the mass conservation of ΦNm\Phi_{N}^{m}, 𝒫NΦΔt,1m\mathcal{P}_{N}\Phi_{\Delta t,1}^{m}, 𝒫NΦΔt,2m\mathcal{P}_{N}\Phi_{\Delta t,2}^{m} and 𝒫NΦh,1m\mathcal{P}_{N}\Phi_{h,1}^{m} (see (5.8), (5.14)), we find that for every m+m\in\mathbb{Z}^{+}, mMm\leq M, it holds

𝒫hΦ^m¯=𝒫hΦNm¯+Δt𝒫h(𝒫NΦΔt,1m)¯+Δt2𝒫h(𝒫NΦΔt,2m)¯+h2𝒫h(𝒫NΦh,1m)¯=𝒫hΦN0¯=ϕ0¯=ϕm¯.\overline{\mathcal{P}_{h}\widehat{\Phi}^{m}}=\overline{\mathcal{P}_{h}\Phi_{N}^{m}}+\Delta t\overline{\mathcal{P}_{h}(\mathcal{P}_{N}\Phi_{\Delta t,1}^{m})}+\Delta t^{2}\overline{\mathcal{P}_{h}(\mathcal{P}_{N}\Phi_{\Delta t,2}^{m})}+h^{2}\overline{\mathcal{P}_{h}(\mathcal{P}_{N}\Phi_{h,1}^{m})}=\overline{\mathcal{P}_{h}\Phi_{N}^{0}}=\overline{\phi^{0}}=\overline{\phi^{m}}. (5.25)

Besides, we infer from (5.23) and (5.25) that the local truncation error τ^n\widehat{\tau}^{n} also fulfills

τ^m¯=0,m+,mM.\overline{\widehat{\tau}^{m}}=0,\quad\forall\,m\in\mathbb{Z}^{+},\ m\leq M. (5.26)

5.2 Error estimate in l(0,T;Hh1)l2(0,T;Hh2)l^{\infty}(0,T;H^{-1}_{h})\cap l^{2}(0,T;H^{2}_{h})

We first derive a preliminary error estimate on the error grid function eme^{m} in the lower order space l(0,T;Hh1)l2(0,T;Hh2)l^{\infty}(0,T;H^{-1}_{h})\cap l^{2}(0,T;H^{2}_{h}). Since em¯=0\overline{e^{m}}=0 for any mm\in\mathbb{N}, mMm\leq M as shown before, the discrete norm 1,h\|\cdot\|_{-1,h} is well defined for eme^{m}.

Proposition 5.1.

Suppose that the exact solution Φ\Phi of the FCH equation (1.1) belongs to the regularity class \mathcal{R} (recall (5.1)) and satisfies the strict separation property (5.2). Then, provided that hh is sufficiently small, and under the linear refinement requirement ΔtC1h\Delta t\leq C_{1}h for some fixed C1>0C_{1}>0, we have

em1,h+[Δtk=1m(ϵ4Δhek22+2λ(λ+ϵpη)ek22)]12C(Δt+h2),\|e^{m}\|_{-1,h}+\left[\Delta t\sum_{k=1}^{m}\Big{(}\epsilon^{4}\|\Delta_{h}e^{k}\|_{2}^{2}+2\lambda(\lambda+\epsilon^{p}\eta)\|e^{k}\|_{2}^{2}\Big{)}\right]^{\frac{1}{2}}\leq C(\Delta t+h^{2}), (5.27)

for any m+m\in\mathbb{Z}^{+} such that tm=mΔtTt_{m}=m\Delta t\leq T, where the constant C>0C>0 is independent of mm, Δt\Delta t and hh.

Proof.

With the aid of the auxiliary profile Φ^\widehat{\Phi} constructed above, we introduce an alternative numerical error function

e^m:=𝒫hΦ^mϕm,m,mM,\widehat{e}^{m}:=\mathcal{P}_{h}\widehat{\Phi}^{m}-\phi^{m},\qquad\forall\,m\in\mathbb{N},\ m\leq M, (5.28)

where ϕm\phi^{m} is the discrete solution to the system (2.18)–(2.19). From Remark 5.3 we see that e^m¯=0\overline{\widehat{e}^{m}}=0. Hence, the discrete norm 1,h\|\cdot\|_{-1,h} is also well defined for the new error grid function e^m\widehat{e}^{m} for all mm\in\mathbb{N}, mMm\leq M.

To proceed, we make the following a priori assumption at the previous time step:

e^n2Δt53+h53.\|\widehat{e}^{n}\|_{2}\leq\Delta t^{\frac{5}{3}}+h^{\frac{5}{3}}. (5.29)

The a priori assumption (5.29) will later be recovered by the optimal rate convergence analysis at the next time step. From (5.29), the linear refinement constraint ΔtC1h\Delta t\leq C_{1}h, and the inverse inequality, we can derive a discrete \|\cdot\|_{\infty} bound for the numerical error function:

e^nCe^n2h32C~h16θ2,\|\widehat{e}^{n}\|_{\infty}\leq\frac{C\|\widehat{e}^{n}\|_{2}}{h^{\frac{3}{2}}}\leq\widetilde{C}h^{\frac{1}{6}}\leq\frac{\theta}{2}, (5.30)

provided that h>0h>0 is sufficiently small (we note that C~>0\widetilde{C}>0 is independent of Δt\Delta t, hh and nn). Here, the constant θ(0,1)\theta\in(0,1) is determined as in (5.21), which is independent of nn. From (5.21) and (5.30), we can deduce that the numerical solution ϕn\phi^{n} also satisfies the strict separation property at the previous time step:

1+θ2(𝒫hΦ^n)i,j,ke^nϕi,j,kn(𝒫hΦ^n)i,j,k+e^n1θ2,1i,j,kN.-1+\frac{\theta}{2}\leq\big{(}\mathcal{P}_{h}\widehat{\Phi}^{n}\big{)}_{i,j,k}-\|\widehat{e}^{n}\|_{\infty}\leq\phi^{n}_{i,j,k}\leq\big{(}\mathcal{P}_{h}\widehat{\Phi}^{n}\big{)}_{i,j,k}+\|\widehat{e}^{n}\|_{\infty}\leq 1-\frac{\theta}{2},\quad 1\leq i,j,k\leq N. (5.31)

For any mm\in\mathbb{N}, mMm\leq M, since e^m¯=0\overline{\widehat{e}^{m}}=0, we find that (Δh)1e^m(-\Delta_{h})^{-1}\widehat{e}^{m} is well defined. This allows us to subtract the numerical scheme (2.18)–(2.19) from (5.23) and then take (discrete) inner product between the resultant and 2(Δh)1e^n+12(-\Delta_{h})^{-1}\widehat{e}^{n+1}. More precisely, we get

1Δt(e^n+11,h2e^n1,h2+e^n+1e^n1,h2)+2ϵ4Δhe^n+122\displaystyle\frac{1}{\Delta t}\big{(}\|\widehat{e}^{n+1}\|_{-1,h}^{2}-\|\widehat{e}^{n}\|_{-1,h}^{2}+\|\widehat{e}^{n+1}-\widehat{e}^{n}\|_{-1,h}^{2}\big{)}+2\epsilon^{4}\|\Delta_{h}\widehat{e}^{n+1}\|_{2}^{2}
+2λ(λ+ϵpη)e^n+122=i=15Ii,\displaystyle\quad+2\lambda(\lambda+\epsilon^{p}\eta)\|\widehat{e}^{n+1}\|_{2}^{2}=\sum_{i=1}^{5}I_{i}, (5.32)

where

I1=2β(Φ^n+1)β(Φ^n+1)β(ϕn+1)β(ϕn+1),e^n+1Ω,\displaystyle I_{1}=-2\left<\beta(\widehat{\Phi}^{n+1})\beta^{\prime}(\widehat{\Phi}^{n+1})-\beta(\phi^{n+1})\beta^{\prime}(\phi^{n+1}),\,\widehat{e}^{n+1}\right>_{\Omega},
I2=2ϵ2ζ=x,y,z(β′′(Φ^n+1)aζ(|DζΦ^n+1|2)β′′(ϕn+1)aζ(|Dζϕn+1|2))\displaystyle I_{2}=-2\epsilon^{2}\left<\sum_{\zeta=x,y,z}\left(\beta^{\prime\prime}(\widehat{\Phi}^{n+1})a_{\zeta}\big{(}|D_{\zeta}\widehat{\Phi}^{n+1}|^{2}\big{)}-\beta^{\prime\prime}(\phi^{n+1})a_{\zeta}\big{(}|D_{\zeta}\phi^{n+1}|^{2}\big{)}\right)\right.
2ζ=x,y,z[dζ(Aζ(β(Φ^n+1))DζΦ^n+1Aζ(β(ϕn+1))Dζϕn+1)],e^n+1Ω,\displaystyle\qquad\qquad\qquad\left.-2\sum_{\zeta=x,y,z}\left[d_{\zeta}\left(A_{\zeta}\big{(}\beta^{\prime}(\widehat{\Phi}^{n+1})\big{)}D_{\zeta}\widehat{\Phi}^{n+1}-A_{\zeta}\big{(}\beta^{\prime}(\phi^{n+1})\big{)}D_{\zeta}\phi^{n+1}\right)\right],\,\widehat{e}^{n+1}\right>_{\Omega},
I3=2λΦ^nβ(Φ^n)λϕnβ(ϕn)+(λ+ϵpη)β(Φ^n)(λ+ϵpη)β(ϕn),e^n+1Ω,\displaystyle I_{3}=2\left<\lambda\widehat{\Phi}^{n}\beta^{\prime}(\widehat{\Phi}^{n})-\lambda\phi^{n}\beta^{\prime}(\phi^{n})+(\lambda+\epsilon^{p}\eta)\beta(\widehat{\Phi}^{n})-(\lambda+\epsilon^{p}\eta)\beta(\phi^{n}),\,\widehat{e}^{n+1}\right>_{\Omega},
I4=2ϵ2(2λ+ϵpη)Δhe^n,e^n+1Ω,\displaystyle I_{4}=-2\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\left<\Delta_{h}\widehat{e}^{n},\,\widehat{e}^{n+1}\right>_{\Omega},
I5=2τ^n+1,e^n+11,h.\displaystyle I_{5}=2\left<\widehat{\tau}^{n+1},\,\widehat{e}^{n+1}\right>_{-1,h}.

Due to the monotonicity of ββ\beta\beta^{\prime} on (1,1)(-1,1), we easily see that I10I_{1}\leq 0. Besides, from the convexity of

β(ϕ),ζ=x,y,zaζ(|Dζϕ|2)Ω,\left<\beta^{\prime}(\phi),\sum_{\zeta=x,y,z}a_{\zeta}\left(\left|D_{\zeta}\phi\right|^{2}\right)\right>_{\Omega},

we can also conclude I20I_{2}\leq 0. Concerning I3I_{3}, we can find some ξ𝒞per\xi\in\mathcal{C}_{\mathrm{per}} with its values ξi,j,k\xi_{i,j,k} staying between (𝒫hΦ^n)i,j,k\big{(}\mathcal{P}_{h}\widehat{\Phi}^{n}\big{)}_{i,j,k} and ϕi,j,kn\phi^{n}_{i,j,k}, 1i,j,kN1\leq i,j,k\leq N, such that

I3=2<[λξβ′′(ξ)+(2λ+ϵpη)β(ξ)]e^n,e^n+1>Ω.I_{3}=2\Big{<}\big{[}\lambda\xi\beta^{\prime\prime}(\xi)+(2\lambda+\epsilon^{p}\eta)\beta^{\prime}(\xi)\big{]}\widehat{e}^{n},\,\widehat{e}^{n+1}\Big{>}_{\Omega}.

Then using (5.21) and (5.31), we obtain

I3C|e^n,e^n+1Ω|2C(e^n22+e^n+122),\displaystyle I_{3}\leq C\left|\left<\widehat{e}^{n},\,\widehat{e}^{n+1}\right>_{\Omega}\right|\leq 2C\big{(}\|\widehat{e}^{n}\|_{2}^{2}+\|\widehat{e}^{n+1}\|_{2}^{2}\big{)},

where C>0C>0 depends on λ\lambda, ϵ\epsilon, η\eta and θ\theta. The remaining two terms I4I_{4} and I5I_{5} can be estimated by simply using the Cauchy-Schwarz inequality such that

I4\displaystyle I_{4} =2ϵ2(2λ+ϵpη)e^n,Δhe^n+1Ωϵ42Δhe^n+122+2(2λ+ϵpη)2e^n22,\displaystyle=-2\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\left<\widehat{e}^{n},\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega}\leq\frac{\epsilon^{4}}{2}\|\Delta_{h}\widehat{e}^{n+1}\|_{2}^{2}+2(2\lambda+\epsilon^{p}\eta)^{2}\|\widehat{e}^{n}\|_{2}^{2},

and

I5e^n+11,h2+τ^n1,h2.I_{5}\leq\|\widehat{e}^{n+1}\|_{-1,h}^{2}+\|\widehat{\tau}^{n}\|_{-1,h}^{2}.

Collecting the above estimates, we infer from (5.2) that

1Δt\displaystyle\frac{1}{\Delta t} (e^n+11,h2e^n1,h2)+32ϵ4Δhe^n+122+2λ(λ+ϵpη)e^n+122\displaystyle\left(\|\widehat{e}^{n+1}\|_{-1,h}^{2}-\|\widehat{e}^{n}\|_{-1,h}^{2}\right)+\frac{3}{2}\epsilon^{4}\|\Delta_{h}\widehat{e}^{n+1}\|_{2}^{2}+2\lambda(\lambda+\epsilon^{p}\eta)\|\widehat{e}^{n+1}\|_{2}^{2}
C^1(e^n22+e^n+122)+e^n+11,h2+τ^n+11,h2,\displaystyle\leq\widehat{C}_{1}\big{(}\|\widehat{e}^{n}\|_{2}^{2}+\|\widehat{e}^{n+1}\|_{2}^{2}\big{)}+\|\widehat{e}^{n+1}\|_{-1,h}^{2}+\|\widehat{\tau}^{n+1}\|_{-1,h}^{2}, (5.33)

where C^1>0\widehat{C}_{1}>0 depends on λ\lambda, ϵ\epsilon, η\eta and θ\theta. On the other hand, using the fact e^n+1¯=0\overline{\widehat{e}^{n+1}}=0, the Hh2H^{2}_{h}-estimate in [66, Proposition 2.2] such that

e^n+1Hh2CΔhe^n+12,\displaystyle\|\widehat{e}^{n+1}\|_{H^{2}_{h}}\leq C\|\Delta_{h}\widehat{e}^{n+1}\|_{2}, (5.34)

and the interpolation inequality, we find

e^n+122CΔhe^n+1223e^n+11,h43.\displaystyle\|\widehat{e}^{n+1}\|_{2}^{2}\leq C\|\Delta_{h}\widehat{e}^{n+1}\|_{2}^{\frac{2}{3}}\|\widehat{e}^{n+1}\|_{-1,h}^{\frac{4}{3}}.

A similar estimate also holds for e^n\widehat{e}^{n}. Applying the above estimates and Young’s inequality, we can eventually write (5.33) as

e^n+11,h2e^n1,h2+Δt[ϵ4Δhe^n+122+2λ(λ+ϵpη)e^n+122]\displaystyle\|\widehat{e}^{n+1}\|_{-1,h}^{2}-\|\widehat{e}^{n}\|_{-1,h}^{2}+\Delta t\Big{[}\epsilon^{4}\|\Delta_{h}\widehat{e}^{n+1}\|_{2}^{2}+2\lambda(\lambda+\epsilon^{p}\eta)\|\widehat{e}^{n+1}\|_{2}^{2}\Big{]}
C^2Δt(e^n+11,h2+e^n12)+ϵ42ΔtΔhe^n22+Δtτ^n+11,h2,\displaystyle\quad\leq\widehat{C}_{2}\Delta t\big{(}\|\widehat{e}^{n+1}\|_{-1,h}^{2}+\|\widehat{e}^{n}\|_{-1}^{2}\big{)}+\frac{\epsilon^{4}}{2}\Delta t\|\Delta_{h}\widehat{e}^{n}\|_{2}^{2}+\Delta t\|\widehat{\tau}^{n+1}\|_{-1,h}^{2},

where C^2>0\widehat{C}_{2}>0 depends on λ\lambda, ϵ\epsilon, η\eta and θ\theta. Thus, for 0<Δt(2C^2)10<\Delta t\leq(2\widehat{C}_{2})^{-1}, using the fact e^0=0\widehat{e}^{0}=0, (5.24) and the discrete Gronwall inequality, we can conclude that

e^n+11,h2+Δtk=1n+1[ϵ4Δhe^k22+2λ(λ+ϵpη)e^k22]C^3(Δt6+h6),\|\widehat{e}^{n+1}\|_{-1,h}^{2}+\Delta t\sum_{k=1}^{n+1}\big{[}\epsilon^{4}\|\Delta_{h}\widehat{e}^{k}\|_{2}^{2}+2\lambda(\lambda+\epsilon^{p}\eta)\|\widehat{e}^{k}\|_{2}^{2}\big{]}\leq\widehat{C}_{3}(\Delta t^{6}+h^{6}), (5.35)

where C^3>0\widehat{C}_{3}>0 depends on λ\lambda, ϵ\epsilon, η\eta and θ\theta and TT.

It remains to recover the a priori assumption (5.29). Obviously, it is valid for n=0n=0. Then we apply an induction argument. Assume that (5.29) holds at the previous time step corresponding to nn\in\mathbb{N}. From (5.35) and the linear refinement constraint ΔtC1h\Delta t\leq C_{1}h, we deduce that

e^n+12h1e^n+11,hC(Δt2+h2)Δt53+h53,\|\widehat{e}^{n+1}\|_{2}\leq h^{-1}\|\widehat{e}^{n+1}\|_{-1,h}\leq C(\Delta t^{2}+h^{2})\leq\Delta t^{\frac{5}{3}}+h^{\frac{5}{3}}, (5.36)

provided that Δt\Delta t and hh are sufficiently small. Hence, (5.29) holds for all nn\in\mathbb{N}.

Now for any mm\in\mathbb{N}, mMm\leq M, recalling the definition of the two error functions (see (5.7) and (5.28)), we have

em1,h\displaystyle\|e^{m}\|_{-1,h} e^m1,h+𝒫hΦNm𝒫hΦ^m1,h\displaystyle\leq\|\widehat{e}^{m}\|_{-1,h}+\|\mathcal{P}_{h}\Phi_{N}^{m}-\mathcal{P}_{h}\widehat{\Phi}^{m}\|_{-1,h}
e^m1,h+Δt𝒫h(𝒫NΦΔt,1m)1,h+Δt2𝒫h(𝒫NΦΔt,2m)1,h+h2𝒫h(𝒫NΦh,1m)1,h,\displaystyle\leq\|\widehat{e}^{m}\|_{-1,h}+\Delta t\|\mathcal{P}_{h}(\mathcal{P}_{N}\Phi_{\Delta t,1}^{m})\|_{-1,h}+\Delta t^{2}\|\mathcal{P}_{h}(\mathcal{P}_{N}\Phi_{\Delta t,2}^{m})\|_{-1,h}+h^{2}\|\mathcal{P}_{h}(\mathcal{P}_{N}\Phi_{h,1}^{m})\|_{-1,h},

and in a similar manner,

em2\displaystyle\|e^{m}\|_{2} e^m2+Δt𝒫h(𝒫NΦΔt,1m)2+Δt2𝒫h(𝒫NΦΔt,2m)2+h2𝒫h(𝒫NΦh,1m)2,\displaystyle\leq\|\widehat{e}^{m}\|_{2}+\Delta t\|\mathcal{P}_{h}(\mathcal{P}_{N}\Phi_{\Delta t,1}^{m})\|_{2}+\Delta t^{2}\|\mathcal{P}_{h}(\mathcal{P}_{N}\Phi_{\Delta t,2}^{m})\|_{2}+h^{2}\|\mathcal{P}_{h}(\mathcal{P}_{N}\Phi_{h,1}^{m})\|_{2},
Δhem2\displaystyle\|\Delta_{h}e^{m}\|_{2} Δhe^m2+ΔtΔh𝒫h(𝒫NΦΔt,1m)2+Δt2Δh𝒫h(𝒫NΦΔt,2m)2+h2Δh𝒫h(𝒫NΦh,1m)2.\displaystyle\leq\|\Delta_{h}\widehat{e}^{m}\|_{2}+\Delta t\|\Delta_{h}\mathcal{P}_{h}(\mathcal{P}_{N}\Phi_{\Delta t,1}^{m})\|_{2}+\Delta t^{2}\|\Delta_{h}\mathcal{P}_{h}(\mathcal{P}_{N}\Phi_{\Delta t,2}^{m})\|_{2}+h^{2}\|\Delta_{h}\mathcal{P}_{h}(\mathcal{P}_{N}\Phi_{h,1}^{m})\|_{2}.

Thanks to the above observations, we can conclude the error estimate (5.27) from (5.35), (5.3) and the uniform boundedness of ΦΔt,1\Phi_{\Delta t,1}, ΦΔt,2\Phi_{\Delta t,2}, Φh,1\Phi_{h,1} and its derivatives. ∎

5.3 Error estimate in l(0,T;Lh2)l2(0,T;Hh3)l^{\infty}(0,T;L_{h}^{2})\cap l^{2}(0,T;H^{3}_{h})

We are now in a position to prove Theorem 5.3.

Proof.

First, from the lower order error estimate (5.35) for the auxiliary numerical error e^n\widehat{e}^{n}, we have obtained a rough Lh2L_{h}^{2} error estimate (5.29), which together with the inverse inequality yields the strict separation property of the numerical solution ϕm\phi^{m} for all mm\in\mathbb{N}, mMm\leq M (see (5.31)).

Next, subtracting the numerical scheme (2.18)–(2.19) from (5.23) and taking (discrete) inner product between the resultant and 2e^n+12\widehat{e}^{n+1}, we get

1Δt(e^n+122e^n22+e^n+1e^n22)+2ϵ4hΔhe^n+122+2λ(λ+ϵpη)he^n+122=i=16Ji,\frac{1}{\Delta t}\left(\|\widehat{e}^{n+1}\|_{2}^{2}-\|\widehat{e}^{n}\|_{2}^{2}+\|\widehat{e}^{n+1}-\widehat{e}^{n}\|_{2}^{2}\right)+2\epsilon^{4}\|\nabla_{h}\Delta_{h}\widehat{e}^{n+1}\|_{2}^{2}+2\lambda(\lambda+\epsilon^{p}\eta)\|\nabla_{h}\widehat{e}^{n+1}\|_{2}^{2}=\sum_{i=1}^{6}J_{i}, (5.37)

where

J1=2(β(Φ^n+1)β(Φ^n+1)β(ϕn+1)β(ϕn+1)),Δhe^n+1Ω,\displaystyle J_{1}=2\left<\left(\beta(\widehat{\Phi}^{n+1})\beta^{\prime}(\widehat{\Phi}^{n+1})-\beta(\phi^{n+1})\beta^{\prime}(\phi^{n+1})\right),\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega},
J2=2ϵ2ζ=x,y,z[β′′(Φ^n+1)aζ(|DζΦ^n+1|2)β′′(ϕn+1)aζ(|Dζϕn+1|2)],Δhe^n+1Ω,\displaystyle J_{2}=2\epsilon^{2}\left<\sum_{\zeta=x,y,z}\left[\beta^{\prime\prime}(\widehat{\Phi}^{n+1})a_{\zeta}\left(|D_{\zeta}\widehat{\Phi}^{n+1}|^{2}\right)-\beta^{\prime\prime}(\phi^{n+1})a_{\zeta}\left(|D_{\zeta}\phi^{n+1}|^{2}\right)\right],\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega},
J3=4ϵ2ζ=x,y,z[dζ(Aζ(β(Φ^n+1))DζΦ^n+1Aζ(β(ϕn+1))Dζϕn+1)],Δhe^n+1Ω,\displaystyle J_{3}=-4\epsilon^{2}\left<\sum_{\zeta=x,y,z}\left[d_{\zeta}\left(A_{\zeta}\big{(}\beta^{\prime}(\widehat{\Phi}^{n+1})\big{)}D_{\zeta}\widehat{\Phi}^{n+1}-A_{\zeta}\big{(}\beta^{\prime}(\phi^{n+1})\big{)}D_{\zeta}\phi^{n+1}\right)\right],\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega},
J4=2(λΦ^nβ(Φ^n)λϕnβ(ϕn)+(λ+ϵpη)β(Φ^n)(λ+ϵpη)β(ϕn)),Δhe^n+1Ω,\displaystyle J_{4}=-2\left<\left(\lambda\widehat{\Phi}^{n}\beta^{\prime}(\widehat{\Phi}^{n})-\lambda\phi^{n}\beta^{\prime}(\phi^{n})+(\lambda+\epsilon^{p}\eta)\beta(\widehat{\Phi}^{n})-(\lambda+\epsilon^{p}\eta)\beta(\phi^{n})\right),\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega},
J5=2ϵ2(2λ+ϵpη)Δhe^n,Δhe^n+1Ω,\displaystyle J_{5}=2\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\left<\Delta_{h}\widehat{e}^{n},\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega},
J6=2τ^n+1,e^n+1Ω.\displaystyle J_{6}=2\left<\widehat{\tau}^{n+1},\,\widehat{e}^{n+1}\right>_{\Omega}.

For J1J_{1}, we can find some ξ1𝒞per\xi_{1}\in\mathcal{C}_{\mathrm{per}} with its values (ξ1)i,j,k(\xi_{1})_{i,j,k} staying between (𝒫hΦ^n+1)i,j,k\big{(}\mathcal{P}_{h}\widehat{\Phi}^{n+1}\big{)}_{i,j,k} and ϕi,j,kn+1\phi^{n+1}_{i,j,k}, 1i,j,kN1\leq i,j,k\leq N, such that

J1\displaystyle J_{1} =2(β(ξ1)2+β(ξ1)β′′(ξ1))e^n+1,Δhe^n+1Ω\displaystyle=2\left<\left(\beta^{\prime}(\xi_{1})^{2}+\beta(\xi_{1})\beta^{\prime\prime}(\xi_{1})\right)\widehat{e}^{n+1},\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega}
Ce^n+12Δhe^n+12\displaystyle\leq C\|\widehat{e}^{n+1}\|_{2}\|\Delta_{h}\widehat{e}^{n+1}\|_{2}
C(Δhe^n+122+e^n+122),\displaystyle\leq C\left(\|\Delta_{h}\widehat{e}^{n+1}\|_{2}^{2}+\|\widehat{e}^{n+1}\|_{2}^{2}\right),

where the constant C>0C>0 only depends on θ\theta. By a similar argument, we can estimate J4J_{4} as

J4\displaystyle J_{4} =2[λ(β(ξ2)+ξ3β′′(ξ2))+(λ+ϵpη)β(ξ2)]e^n,Δhe^n+1Ω\displaystyle=-2\left<\big{[}\lambda(\beta^{\prime}(\xi_{2})+\xi_{3}\beta^{\prime\prime}(\xi_{2}))+(\lambda+\epsilon^{p}\eta)\beta^{\prime}(\xi_{2})\big{]}\widehat{e}^{n},\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega}
Ce^n2Δhe^n+12\displaystyle\leq C\|\widehat{e}^{n}\|_{2}\|\Delta_{h}\widehat{e}^{n+1}\|_{2}
C(Δhe^n+122+e^n22),\displaystyle\leq C\left(\|\Delta_{h}\widehat{e}^{n+1}\|_{2}^{2}+\|\widehat{e}^{n}\|_{2}^{2}\right),

where the constant C>0C>0 only depends on λ\lambda, η\eta, ϵ\epsilon and θ\theta.

Next, we handle J2J_{2} by making the following decomposition:

J2\displaystyle J_{2} =2ϵ2(ζ=x,y,z(β′′(Φ^n+1)β′′(ϕn+1))aζ(|DζΦ^n+1|2),Δhe^n+1Ω\displaystyle=2\epsilon^{2}\left(\left<\sum_{\zeta=x,y,z}\left(\beta^{\prime\prime}(\widehat{\Phi}^{n+1})-\beta^{\prime\prime}(\phi^{n+1})\right)a_{\zeta}\left(|D_{\zeta}\widehat{\Phi}^{n+1}|^{2}\right),\,\Delta_{h}\hat{e}^{n+1}\right>_{\Omega}\right.
+ζ=x,y,zβ′′(ϕn+1)aζ(|DζΦ^n+1|2|Dζϕn+1|2),Δhe^n+1Ω)\displaystyle\qquad+\left.\left<\sum_{\zeta=x,y,z}\beta^{\prime\prime}(\phi^{n+1})a_{\zeta}\left(|D_{\zeta}\widehat{\Phi}^{n+1}|^{2}-|D_{\zeta}\phi^{n+1}|^{2}\right),\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega}\right)
=:2ϵ2(J2a+J2b).\displaystyle=:2\epsilon^{2}(J_{2a}+J_{2b}).

Again, we can find some ξ3𝒞per\xi_{3}\in\mathcal{C}_{\mathrm{per}} with its values (ξ3)i,j,k(\xi_{3})_{i,j,k} staying between (𝒫hΦ^n+1)i,j,k\big{(}\mathcal{P}_{h}\widehat{\Phi}^{n+1}\big{)}_{i,j,k} and ϕi,j,kn+1\phi^{n+1}_{i,j,k}, 1i,j,kN1\leq i,j,k\leq N, such that

J2a\displaystyle J_{2a} =ζ=x,y,zβ′′′(ξ3)e^n+1aζ(|DζΦ^n+1|2),Δhe^n+1Ω\displaystyle=\left<\sum_{\zeta=x,y,z}\beta^{\prime\prime\prime}(\xi_{3})\widehat{e}^{n+1}a_{\zeta}\left(|D_{\zeta}\widehat{\Phi}^{n+1}|^{2}\right),\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega}
Cζ=x,y,zaζ(|DζΦ^n+1|2)e^n+12Δhe^n+12\displaystyle\leq C\left\|\sum_{\zeta=x,y,z}a_{\zeta}\left(|D_{\zeta}\widehat{\Phi}^{n+1}|^{2}\right)\right\|_{\infty}\|\widehat{e}^{n+1}\|_{2}\|\Delta_{h}\widehat{e}^{n+1}\|_{2}
Ce^n+12Δhe^n+12,\displaystyle\leq C\|\widehat{e}^{n+1}\|_{2}\|\Delta_{h}\widehat{e}^{n+1}\|_{2},

where C>0C>0 depends on θ\theta and norms of Φ\Phi. For the second term, we infer from (5.31), Hölder’s inequality and the uniform Hh2H_{h}^{2}-bound (4.2) that

J2b\displaystyle J_{2b} C|ζ=x,y,zaζ(|Dζ(Φ^n+1+ϕn+1)||Dζe^n+1|),Δhe^n+1Ω|\displaystyle\leq C\left|\left<\sum_{\zeta=x,y,z}a_{\zeta}\left(|D_{\zeta}(\widehat{\Phi}^{n+1}+\phi^{n+1})||D_{\zeta}\widehat{e}^{n+1}|\right),\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega}\right|
Cζ=x,y,zaζ(|Dζ(Φ^n+1+ϕn+1)|)Lh4ζ=x,y,z(aζ|Dζe^n+1|)Lh4Δhe^n+12\displaystyle\leq C\left\|\sum_{\zeta=x,y,z}a_{\zeta}\left(|D_{\zeta}(\widehat{\Phi}^{n+1}+\phi^{n+1})|\right)\right\|_{L_{h}^{4}}\left\|\sum_{\zeta=x,y,z}\left(a_{\zeta}|D_{\zeta}\widehat{e}^{n+1}|\right)\right\|_{L_{h}^{4}}\left\|\Delta_{h}\widehat{e}^{n+1}\right\|_{2}
C(hΦ^n+1Lh4+hϕn+1Lh4)he^n+1Lh4Δhe^n+12\displaystyle\leq C\left(\|\nabla_{h}\widehat{\Phi}^{n+1}\|_{L_{h}^{4}}+\|\nabla_{h}\phi^{n+1}\|_{L_{h}^{4}}\right)\left\|\nabla_{h}\widehat{e}^{n+1}\right\|_{L_{h}^{4}}\left\|\Delta_{h}\widehat{e}^{n+1}\right\|_{2}
C(Φ^n+1Hh2+ϕn+1Hh2)e^n+1Hh2Δhe^n+12\displaystyle\leq C\left(\|\widehat{\Phi}^{n+1}\|_{H^{2}_{h}}+\|\phi^{n+1}\|_{H^{2}_{h}}\right)\|\widehat{e}^{n+1}\|_{H^{2}_{h}}\|\Delta_{h}\widehat{e}^{n+1}\|_{2}
Ce^n+1Hh2Δhe^n+12.\displaystyle\leq C\|\widehat{e}^{n+1}\|_{H^{2}_{h}}\|\Delta_{h}\widehat{e}^{n+1}\|_{2}.

Here, we have used the estimate ζ=x,y,zaζ|Dζf|Lh4ChfLh4\left\|\sum_{\zeta=x,y,z}a_{\zeta}|D_{\zeta}f|\right\|_{L_{h}^{4}}\leq C\left\|\nabla_{h}f\right\|_{L^{4}_{h}} for all f𝒞perf\in\mathcal{C}_{\mathrm{per}} and discrete interpolation inequalities. Combining the estimates for J2aJ_{2a} and I2bI_{2b}, we deduce from (5.34) that

J2Ce^n+1Hh2Δhe^n+12CΔhe^n+122.J_{2}\leq C\|\widehat{e}^{n+1}\|_{H^{2}_{h}}\|\Delta_{h}\widehat{e}^{n+1}\|_{2}\leq C\|\Delta_{h}\widehat{e}^{n+1}\|_{2}^{2}.

The term J3J_{3} can be treated in a similar manner. Indeed, we have

J3\displaystyle J_{3} 4ϵ2(|ζ=x,y,zdζ(Aζ(β(Φ^n+1)β(ϕn+1))DζΦ^n+1),Δhe^n+1Ω|\displaystyle\leq 4\epsilon^{2}\left(\left|\left<\sum_{\zeta=x,y,z}d_{\zeta}\left(A_{\zeta}\big{(}\beta^{\prime}(\widehat{\Phi}^{n+1})-\beta^{\prime}(\phi^{n+1})\big{)}D_{\zeta}\widehat{\Phi}^{n+1}\right),\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega}\right|\right.
+|ζ=x,y,zdζ(Aζβ(ϕn+1)Dζe^n+1),Δhe^n+1Ω|)\displaystyle\qquad+\left.\left|\left<\sum_{\zeta=x,y,z}d_{\zeta}\big{(}A_{\zeta}\beta^{\prime}(\phi^{n+1})D_{\zeta}\widehat{e}^{n+1}\big{)},\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega}\right|\right)
=:4ϵ2(J3a+J3b).\displaystyle=:4\epsilon^{2}\left(J_{3a}+J_{3b}\right).

Concerning J3aJ_{3a}, there exists some ξ4𝒞per\xi_{4}\in\mathcal{C}_{\mathrm{per}} with its values (ξ4)i,j,k(\xi_{4})_{i,j,k} staying between (𝒫hΦ^n+1)i,j,k\big{(}\mathcal{P}_{h}\widehat{\Phi}^{n+1}\big{)}_{i,j,k} and ϕi,j,kn+1\phi^{n+1}_{i,j,k}, 1i,j,kN1\leq i,j,k\leq N, such that

J3a\displaystyle J_{3a} =|ζ=x,y,zdζ(Aζ(β′′(ξ4)e^n+1)DζΦ^n+1),Δhe^n+1Ω|\displaystyle=\left|\left<\sum_{\zeta=x,y,z}d_{\zeta}\left(A_{\zeta}\big{(}\beta^{\prime\prime}(\xi_{4})\widehat{e}^{n+1}\big{)}D_{\zeta}\widehat{\Phi}^{n+1}\right),\,\Delta_{h}\widehat{e}^{n+1}\right>_{\Omega}\right|
Cζ=x,y,zβ′′(ξ4)DζΦ^n+1e^n+12hΔhe^n+12\displaystyle\leq C\sum_{\zeta=x,y,z}\|\beta^{\prime\prime}(\xi_{4})\|_{\infty}\|D_{\zeta}\widehat{\Phi}^{n+1}\|_{\infty}\|\widehat{e}^{n+1}\|_{2}\|\nabla_{h}\Delta_{h}\widehat{e}^{n+1}\|_{2}
Ce^n+12hΔhe^n+12,\displaystyle\leq C\left\|\widehat{e}^{n+1}\right\|_{2}\left\|\nabla_{h}\Delta_{h}\widehat{e}^{n+1}\right\|_{2},

where we have used Lemma 2.1 and Hölder’s inequality. Similarly, for J3bJ_{3b}, we have

J3bChe^n+12hΔhe^n+12.J_{3b}\leq C\left\|\nabla_{h}\widehat{e}^{n+1}\right\|_{2}\left\|\nabla_{h}\Delta_{h}\widehat{e}^{n+1}\right\|_{2}.

As a consequence, it holds

J3\displaystyle J_{3} Ce^n+1Hh1hΔhe^n+12\displaystyle\leq C\left\|\widehat{e}^{n+1}\right\|_{H^{1}_{h}}\left\|\nabla_{h}\Delta_{h}\widehat{e}^{n+1}\right\|_{2}
ϵ4hΔhe^n+122+Ce^n+1Hh12\displaystyle\leq\epsilon^{4}\|\nabla_{h}\Delta_{h}\widehat{e}^{n+1}\|_{2}^{2}+C\left\|\widehat{e}^{n+1}\right\|_{H^{1}_{h}}^{2}
ϵ4hΔhe^n+122+CΔhe^n+122.\displaystyle\leq\epsilon^{4}\|\nabla_{h}\Delta_{h}\widehat{e}^{n+1}\|_{2}^{2}+C\left\|\Delta_{h}\widehat{e}^{n+1}\right\|_{2}^{2}.

For the remaining two terms, an application of Cauchy-Schwarz inequality yields that

J5\displaystyle J_{5} 2ϵ2(2λ+ϵpη)Δhe^n2Δhe^n+12ϵ2(2λ+ϵpη)(Δhe^n+122+Δhe^n22),\displaystyle\leq 2\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\|\Delta_{h}\widehat{e}^{n}\|_{2}\|\Delta_{h}\widehat{e}^{n+1}\|_{2}\leq\epsilon^{2}(2\lambda+\epsilon^{p}\eta)\big{(}\|\Delta_{h}\widehat{e}^{n+1}\|_{2}^{2}+\|\Delta_{h}\widehat{e}^{n}\|_{2}^{2}\big{)},
J6\displaystyle J_{6} e^n+122+τ^n+122.\displaystyle\leq\|\widehat{e}^{n+1}\|_{2}^{2}+\|\widehat{\tau}^{n+1}\|_{2}^{2}.

Combining the above estimates, we can deduce from (5.34) and (5.37) that

1Δt(e^n+122e^n22)+ϵ4hΔhe^n+122+2λ(λ+ϵpη)he^n+122\displaystyle\frac{1}{\Delta t}\left(\|\widehat{e}^{n+1}\|_{2}^{2}-\|\widehat{e}^{n}\|_{2}^{2}\right)+\epsilon^{4}\|\nabla_{h}\Delta_{h}\widehat{e}^{n+1}\|_{2}^{2}+2\lambda(\lambda+\epsilon^{p}\eta)\|\nabla_{h}\widehat{e}^{n+1}\|_{2}^{2}
C(Δhe^n+122+Δhe^n22)+τ^n+122,\displaystyle\qquad\leq C\left(\|\Delta_{h}\widehat{e}^{n+1}\|_{2}^{2}+\|\Delta_{h}\widehat{e}^{n}\|_{2}^{2}\right)+\|\widehat{\tau}^{n+1}\|_{2}^{2}, (5.38)

which together with the fact e^0=0\widehat{e}^{0}=0 leads to the following inequality

e^n+122+Δtk=1n+1[ϵ4hΔhe^k22+2λ(λ+ϵpη)he^k22]\displaystyle\|\widehat{e}^{n+1}\|_{2}^{2}+\Delta t\sum_{k=1}^{n+1}\left[\epsilon^{4}\big{\|}\nabla_{h}\Delta_{h}\widehat{e}^{k}\big{\|}_{2}^{2}+2\lambda(\lambda+\epsilon^{p}\eta)\|\nabla_{h}\widehat{e}^{k}\|_{2}^{2}\right]
CΔtk=1n+1Δhe^k22+Δtk=0nτ^k+122.\displaystyle\quad\leq C\Delta t\sum_{k=1}^{n+1}\|\Delta_{h}\widehat{e}^{k}\|_{2}^{2}+\Delta t\sum_{k=0}^{n}\big{\|}\widehat{\tau}^{k+1}\big{\|}_{2}^{2}. (5.39)

Thanks to the truncation error (5.24) and the lower order error estimate (5.35), we deduce from (5.39) that

e^n+122+Δtk=1n+1[ϵ4hΔhe^k22+2λ(λ+ϵpη)he^k22]C(Δt6+h6).\left\|\widehat{e}^{n+1}\right\|_{2}^{2}+\Delta t\sum_{k=1}^{n+1}\left[\epsilon^{4}\big{\|}\nabla_{h}\Delta_{h}\widehat{e}^{k}\big{\|}_{2}^{2}+2\lambda(\lambda+\epsilon^{p}\eta)\|\nabla_{h}\widehat{e}^{k}\|_{2}^{2}\right]\leq C(\Delta t^{6}+h^{6}). (5.40)

Finally, by a comparison between the two error functions e^n+1\widehat{e}^{n+1} and en+1e^{n+1} (using (5.7), (5.11) and (5.28)), from the projection estimate (5.3), the uniform boundedness of ΦΔt,1\Phi_{\Delta t,1}, ΦΔt,2\Phi_{\Delta t,2}, Φh,1\Phi_{h,1}, the estimates (5.35), (5.40), we can achieve the conclusion (5.9).

This completes the proof of Theorem 5.3. ∎

6 Numerical Experiments

In this section, we implement a preconditioned steepest descent (PSD) algorithm to solve the fully discrete numerical scheme (2.18)–(2.19), following the framework proposed in [67]. This algorithm has been applied to the FCH equation (1.1) with a regular potential (1.6) in [45]. Further applications can be found in [68] and the references cited therein.

Observe that our scheme (2.18)–(2.19) can be recast as a minimization problem of the discrete energy (2.14). This is equivalent to solve the zero point of the discrete variation of (2.14):

𝒩h(ϕ)=fn,\mathcal{N}_{h}(\phi)=f^{n},

where

𝒩h(ϕ)=ϕΔtΔhδϕEhc(ϕ)andfn=ϕnΔt+ΔhδϕEhe(ϕn).\mathcal{N}_{h}(\phi)=\frac{\phi}{\Delta t}-\Delta_{h}\delta_{\phi}E_{h}^{c}(\phi)\quad\text{and}\quad f^{n}=\frac{\phi^{n}}{\Delta t}+\Delta_{h}\delta_{\phi}E_{h}^{e}(\phi^{n}).

The essential idea of the PSD solver is to use a linearized version of the nonlinear operator as a pre-conditioner. Here, the preconditioner h:𝒞̊per𝒞̊per\mathcal{L}_{h}:\mathring{\mathcal{C}}_{\mathrm{per}}\rightarrow\mathring{\mathcal{C}}_{\mathrm{per}} is defined as

h:=1Δtϵ4Δh3+ϵ2θ1Δh2(λ2+λϵpη+θ2)Δh,\mathcal{L}_{h}:=\begin{aligned} &\frac{1}{\Delta t}-\epsilon^{4}\Delta_{h}^{3}+\epsilon^{2}\theta_{1}\Delta_{h}^{2}-(\lambda^{2}+\lambda\epsilon^{p}\eta+\theta_{2})\Delta_{h},\end{aligned}

where θ1\theta_{1}, θ2>0\theta_{2}>0 are two parameters that can be adjusted. Obviously, h\mathcal{L}_{h} is a positive, symmetric operator. In particular, this metric can be used to find an appropriate search direction for the steepest descent solver. Given the current iterate ϕn\phi^{n}, we obtain the next iterate direction by finding a dn𝒞perd_{n}\in\mathcal{C}_{\mathrm{per}} such that

h[dn]=rnrn¯,rn¯:=fn𝒩h(ϕn),\mathcal{L}_{h}[d_{n}]=r_{n}-\overline{r_{n}},\qquad\overline{r_{n}}:=f^{n}-\mathcal{N}_{h}(\phi^{n}),

where rnr_{n} is the nonlinear residual of the nthn^{\mathrm{th}} iterate ϕn\phi^{n}. Using this linear operator, we then solve the equation by the Fast Fourier Transform (FFT). Consequently, the next iterate is obtained as

ϕn+1:=ϕn+αndn,\phi^{n+1}:=\phi^{n}+\alpha_{n}d_{n},

where αn\alpha_{n}\in\mathbb{R} is the unique solution to the steepest descent line minimization problem

αn:=argminαEh(ϕn+αdn)=argzeroα𝒩h(ϕn+αdn)fn,dnΩ.\alpha_{n}:=\mathop{\rm{argmin}}\limits_{\alpha\in\mathbb{R}}E_{h}(\phi^{n}+\alpha d_{n})=\mathop{\rm{argzero}}\limits_{\alpha\in\mathbb{R}}\left<\mathcal{N}_{h}(\phi^{n}+\alpha d_{n})-f^{n},d_{n}\right>_{\Omega}.

6.1 Convergence test

In this subsection we carry out an accuracy check for the numerical scheme (2.18)–(2.19). For simplicity, the computational domain is assumed to be the unit square Ω=(0,1)2\Omega=(0,1)^{2}, and the phase variable is given as

Φ(x,y,t)=1πsin(2πx)cos(2πy)cos(t).\Phi(x,y,t)=\frac{1}{\pi}\sin(2\pi x)\cos(2\pi y)\cos(t).

In particular, we see that 1+Φ1+\Phi and 1Φ1-\Phi stays positive at a pointwise level, so no singularity occurs during the computation. In order to make Φ\Phi become the solution of the PDE system (1.1), we have to add an artificial, time dependent forcing term, then we can continue to finish the convergence test.

We first give the parameters as ϵ=0.5\epsilon=0.5, η=1\eta=1, λ=3\lambda=3 and p=2p=2. We compute the Lh2L_{h}^{2} numerical errors with grid size N=16, 32, 64, 128, 256, 512N=16,\,32,\,64,\,128,\,256,\,512. To verify that our numerical scheme has first order in time and second order in space accuracy, we set the time step size Δt=16h2\Delta t=16h^{2} and Δt=h\Delta t=h respectively, and plot ln|e|\ln|e| versus lnN\ln N to demonstrate the temporal convergence order. In the first case Δt=16h2\Delta t=16h^{2}, the expected temporal numerical accuracy assumption e=C(Δt+h2)e=C(\Delta t+h^{2}) indicates that ln|e|=lnC2lnN\ln|e|=\ln C-2\ln N. This is consistent with the red lines shown in Figure 1, with its slope approximately being 2.0498-2.0498. In the second case Δt=h\Delta t=h, the numerical error gradually approach the line with slope 1-1, as depicted with the purple dotted line.

Refer to caption
Figure 1: The discrete Lh2L_{h}^{2} numerical error versus spatial resolution NN for N=16, 32, 64, 128, 256, 512N=16,\,32,\,64,\,128,\,256,\,512. The blue dots represent the numerical error in the case Δt=16h2\Delta t=16h^{2}. The data are roughly on curves CN2CN^{-2}, as is demonstrated by the red fitted line. Besides, the orange dots represent the numerical error in the case Δt=h\Delta t=h. The data have an asymptote with a slope 1-1 (plotted by the purple dotted line), which verifies the first order in time accuracy.

6.2 Pearling bifurcation and meandering instability

6.2.1 Adaptive time adjust strategy

Hereafter, we apply an adaptive time strategy to perform our simulation. We set an upper bound and lower bound for both the energy decay rate and the phase change rate. If either of the two factors changes rapidly, that is, exceeding the upper bound, we shorten the time step, and if both factors change too slow, we increase the time step. The maximal time step is given as 2e32e-3, and the upper bound and lower bound are set to be 1e11e-1 and 1e31e-3, respectively.

6.2.2 The pearling bifurcation and sensitivity of the initial value

The pearling bifurcation is an important physical phenomenon that has been investigated in [12, 69, 13]. Several numerical simulations have been implemented in recent works [49, 48, 46], always with a regular potential. In our simulation, the following initial condition will be used:

ϕ(x,y,0)=1.8/cosh((x0.5)2+(y0.5)20.42ϵ)0.9,\phi(x,y,0)=1.8\Big{/}\cosh\left(\frac{\sqrt{(x-0.5)^{2}+(y-0.5)^{2}}-0.42}{\ell\epsilon}\right)-0.9, (6.1)

where the parameter >0\ell>0 controls the interface width of the initial datum. Here we set h=1/256h=1/256, ϵ=0.03\epsilon=0.03, η=4\eta=4, p=1p=1 and λ=ln(19)/0.9\lambda=\ln(19)/0.9 so that the minima ±r\pm\,r_{*} of FF are given with r=0.9r_{*}=0.9. As a consequence, Figure 2, Figure 3 and Figure 4 show the time snapshots of simulations under different initial thickness of the ring. The pearl bifurcation appears when =0.35\ell=0.35 and =0.39\ell=0.39, but fails to exist when =0.5\ell=0.5. From this numerical experiment we can see that, the morphological structure of interfaces is highly sensitive to the initial datum. A slight change may lead to different intermediate states and long-time equilibria. The energy plot in Figure 5 shows a fast exponential decay at the beginning of the simulation, and when the pearl structure appears, the energy changes fast as well. In Figure 6 we observe that in the above three cases for pearling bifurcation, the FCH energy, the Cahn-Hilliard energy ECHE_{\mathrm{CH}} and the phase-field Willmore (PFW) energy decrease at the same time.

Refer to caption
(a) t=0t=0
Refer to caption
(b) t=0.3t=0.3
Refer to caption
(c) t=0.5t=0.5
Refer to caption
(d) t=0.8t=0.8
Refer to caption
(e) t=10t=10
Figure 2: Snapshots with initial condition (6.1) when =0.35\ell=0.35. The ring splits near t=0.3t=0.3 and converges to 4 circles.
Refer to caption
(a) t=0t=0
Refer to caption
(b) t=0.3t=0.3
Refer to caption
(c) t=0.5t=0.5
Refer to caption
(d) t=0.8t=0.8
Refer to caption
(e) t=10t=10
Figure 3: Snapshots with initial condition (6.1) when =0.39\ell=0.39. The pearl bifurcation appears near t=0.8t=0.8, and converges to 8 circles.
Refer to caption
(a) t=0t=0
Refer to caption
(b) t=0.3t=0.3
Refer to caption
(c) t=0.5t=0.5
Refer to caption
(d) t=0.8t=0.8
Refer to caption
(e) t=10t=10
Figure 4: Snapshots with initial condition (6.1) when =0.5\ell=0.5. The ring doesn’t split but only shrink and thicken. The pearl structure fails to exist the whole time.
\begin{overpic}[scale={.45}]{Figs/pearl/energy.eps} \put(48.0,30.0){\includegraphics[scale={.2},trim=30.11249pt 20.075pt 30.11249pt 20.075pt,clip]{Figs/pearl/energy2.eps}} \end{overpic}
Figure 5: Loglog plot of the FCH energy in three different cases. Four time ticks of the snapshots in Figure 2-4 are marked by the dotted lines. All cases encounter a fast exponential energy decay at the beginning of the simulation. When the pearl bifurcation appears, the energy also changes rapidly.
\begin{overpic}[scale={.4}]{Figs/pearl/total.eps} \put(91.0,12.0){\includegraphics[scale={.3},trim=0.0pt 0.0pt 210.78749pt 0.0pt,clip]{Figs/pearl/total2.eps}} \end{overpic}
Figure 6: Loglog plot of the FCH energy, the minus Cahn-Hilliard energy and the phase-field Willmore (PFW) energy in three different cases. We observe that both the PFW energy and the Cahn-Hilliard energy decrease during time evolution for pearling bifurcation.

6.2.3 The meandering instability

The meandering instability is another interesting phenomenon investigated in the literature, see e.g., [12]. Some numerical tests can be found in [45, 46] with a regular potential. To proceed with our simulation, we consider the following initial condition:

ϕ(x,y,0)={0.9,ify>0.5sin(4πx12)+6.4,0.9,ify<0.5sin(4πx15)5.6,0.9,otherwise.\phi(x,y,0)=\left\{\begin{aligned} &-0.9,\quad\mbox{if}\ y>0.5\sin\left(\frac{4\pi x}{12}\right)+6.4,\\ &-0.9,\quad\mbox{if}\ y<0.5\sin\left(\frac{4\pi x}{15}\right)-5.6,\\ &0.9,\qquad\ \mbox{otherwise}.\end{aligned}\right. (6.2)

Besides, we set η=10\eta=10, ϵ=0.01\epsilon=0.01, p=1p=1, h=1/256h=1/256 and λ=ln(19)/0.9\lambda=\ln(19)/0.9. Figure 7 shows the time snapshots of the meandering instability, and Figure LABEL:meanderenergy1 presents the FCH and Cahn-Hilliard energy change of this experiment. The FCH energy still decreases over time, however, the Cahn-Hilliard energy increases after the initial decay. This reflects the balancing between the two parts (PFW vs. Cahn-Hilliard) in the FCH energy.

Refer to caption
(a) t=0t=0
Refer to caption
(b) t=1t=1
Refer to caption
(c) t=5t=5
Refer to caption
(d) t=15t=15
Refer to caption
(e) t=30t=30
Refer to caption
(f) t=50t=50
Refer to caption
(g) t=60t=60
Refer to caption
(h) t=100t=100
Figure 7: Snapshots of the meandering instability with initial condition (6.2). Parameters are set to be as η=10,ϵ=0.01,p=1\eta=10,\ \epsilon=0.01,\ p=1, h=1/256h=1/256 and λ=ln(19)/0.9\lambda=\ln(19)/0.9.

6.3 Phase separation

In this subsection we simulate the phase separation process, namely, the spinodal decomposition of a binary mixture with possible amphiphilic structure. We start with the following random initial condition (cf. [45]):

ϕ(x,y,0)=0.5+0.01(2r1),\phi(x,y,0)=0.5+0.01(2r-1), (6.3)

where rr are uniformly distributed random numbers in [0,1][0,1], and give the parameters as ϵ=0.008\epsilon=0.008, η=8\eta=8, p=1p=1, h=1/256h=1/256 and λ=ln(19)/0.9\lambda=\ln(19)/0.9.

Snapshots of the phase separation process can be found in Figure 9. The minimum value and maximum value of the phase variable stays between 0.8852-0.8852 and 0.90350.9035 along time evolution, which keep a safe distance to the pure states 1-1 and 11. This is consistent with the strict separation property proved in this paper (for the numerical scheme) and in [42] (for the partial differential equation). Meanwhile, the average mass has a 101210^{-12} error after 10510^{5} iteration, which comes from possible cancellation errors when rounding off values at machine precision, as is shown in Figure 11. The energy plot show a similar property to that for the previous simulation on meandering instability, in which the Cahn-Hilliard energy also increases after the initial decay. Figure 10 shows some special structures that we capture in the simulation, and these can be also found in the experiment of [70].

Refer to caption
(a) t=0t=0
Refer to caption
(b) t=0.01t=0.01
Refer to caption
(c) t=0.1t=0.1
Refer to caption
(d) t=1t=1
Refer to caption
(e) t=10t=10
Refer to caption
(f) t=50t=50
Refer to caption
(g) t=100t=100
Refer to caption
(h) t=500t=500
Figure 9: Time snapshots of phase separation with the random initial datum (6.3).
Refer to caption
(a) Y-juntion
Refer to caption
(b) cylinder
Refer to caption
(c) sphere
Refer to caption
(d) cylindrical loop
Figure 10: Special structures like Y-junction, cylinder, spheres and cylindrical loops shown in the phase separation, which is consistent with Figure 1 and Figure 3 in [70].
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 11: Minimum value (top left), maximum value (top right), averaged mass (bottom left) and free energy (bottom right) plot in the long time simulation. These four pictures respectively show that our discrete numerical scheme inherits the properties of strict separation from pure states, mass conservation and energy dissipation that have been proved for the continuous problem.

7 Concluding Remarks

In this paper, we present a first order semi-implicit finite difference scheme for the functionalized Cahn-Hilliard equation with a logarithmic Flory-Huggins type potential. The numerical scheme is derived based on the convex splitting technique. The non-convex and non-concave term Ωβ(ϕ)Δϕdx-\int_{\Omega}\beta(\phi)\Delta\phi\,\mathrm{d}x in the FCH energy is handled by using integration by parts under periodic boundary conditions. Unique solvability, unconditional energy stability and mass conservation property of the numerical scheme are theoretically justified. Moreover, we establish the positivity-preserving property, i.e., the phase function ϕ\phi stays in (1,1)(-1,1) at a point-wise sense, with the aid of the singular nature of the logarithmic term β(ϕ)\beta(\phi). To the best of our knowledge, our numerical scheme is the first one that combines these four important theoretic properties for the FCH equation with a singular potential. Next, we perform an optimal rate convergence analysis and obtain error estimates in the higher order space l(0,T;Lh2)l2(0,T;Hh3)l^{\infty}(0,T;L_{h}^{2})\cap l^{2}(0,T;H_{h}^{3}) under a linear refinement requirement ΔtC1h\Delta t\leq C_{1}h. To achieve the goal, we choose to perform a higher order asymptotic expansion up to third order accuracy in time and space, with a careful linearization technique. We first derive an error estimate in the lower order space l(0,T;Hh1)l2(0,T;Hh2)l^{\infty}(0,T;H_{h}^{-1})\cap l^{2}(0,T;H^{2}_{h}) and obtain a rough error estimate in Lh2L_{h}^{2}. Then the Lh2L_{h}^{2} error estimate combined with the inverse inequality yields a uniform LhL_{h}^{\infty} bound on the error function, which implies the strict separation from pure states ±1\pm 1 for the numerical solutions. This crucial fact enables us to achieve the refined Lh2L_{h}^{2} estimate by using the energy method. In the last part of this paper, we present some numerical experiments in the two dimensional case with the PSD solver, which demonstrate the accuracy and robustness of the discrete scheme. Pearling bifurcation, meandering instability and spinodal decomposition are observed via the numerical simulation. The energy stability, mass conservation and the positivity preserving property are also verified numerically.

Acknowledgements

W. Chen was partially supported by NSFC 12241101 and NSFC 12071090. H. Wu was partially supported by NSFC 12071084 and the Shanghai Center for Mathematical Sciences at Fudan University. W. Chen and H. Wu are members of the Key Laboratory of Mathematics for Nonlinear Sciences (Fudan University), Ministry of Education of China.

References

  • [1] J. W. Cahn, J. E. Hilliard, Free energy of a nonuniform system. I. interfacial free energy, The Journal of Chemical Physics 28 (2) (1958) 258–267.
  • [2] S. Dai, K. Promislow, Geometric evolution of bilayers under the functionalized Cahn-Hilliard equation, Proc. Roy. Soc. A 469 (2013) 20120505, 20 pp.
  • [3] Q. Du, C. Liu, X. Wang, A phase field approach in the numerical study of the elastic bending energy for vesicle membranes, J. Comput. Phys. 198 (2) (2004) 450–468.
  • [4] Q. Du, C. Liu, R. Ryham, X. Wang, A phase field formulation of the Willmore problem, Nonlinearity 18 (3) (2005) 1249–1267.
  • [5] S. Torabi, J. Lowengrub, A. Voigt, S. M. Wise, A new phase-field model for strongly anisotropic systems, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 465 (2105) (2009) 1337–1359, with supplementary material available online.
  • [6] S. M. Wise, J. Kim, J. Lowengrub, Solving the regularized, strongly anisotropic Cahn-Hilliard equation by an adaptive nonlinear multigrid method, J. Comput. Phys. 226 (1) (2007) 414–446.
  • [7] F. Chen, J. Shen, Efficient energy stable schemes with spectral discretization in space for anisotropic Cahn-Hilliard systems, Commun. Comput. Phys. 13 (5) (2013) 1189–1208.
  • [8] G. Gompper, M. Schick, Correlation between structural and interfacial properties of amphiphilic systems, Phys. Rev. Lett. 65 (9) (1990) 1116–1119.
  • [9] K. Promislow, B. Wetton, PEM fuel cells: a mathematical overview, SIAM J. Appl. Math. 70 (2) (2009) 369–409.
  • [10] S. Dai, K. Promislow, Competitive geometric evolution of amphiphilic interfaces, SIAM J. Math. Anal. 47 (1) (2015) 347–380.
  • [11] N. Gavish, G. Hayrapetyan, K. Promislow, L. Yang, Curvature driven flow of bi-layer interfaces, Phys. D 240 (7) (2011) 675–693.
  • [12] A. Doelman, G. Hayrapetyan, K. Promislow, B. Wetton, Meander and pearling of single-curvature bilayer interfaces in the functionalized Cahn-Hilliard equation, SIAM J. Math. Anal. 46 (6) (2014) 3640–3677.
  • [13] K. Promislow, Q. Wu, Existence of pearled patterns in the planar functionalized Cahn-Hilliard equation, J. Differential Equations 259 (7) (2015) 3298–3343.
  • [14] J. Jones, Development of a fast and accurate time stepping scheme for the functionalized Cahn-Hilliard equation and application to a graphics processing unit, ProQuest LLC, Ann Arbor, MI, 2013, thesis (Ph.D.)–Michigan State University.
  • [15] N. Kraitzman, K. Promislow, Pearling bifurcations in the strong functionalized Cahn-Hilliard free energy, SIAM J. Math. Anal. 50 (3) (2018) 3395–3426.
  • [16] K. Promislow, Q. Wu, Existence, bifurcation, and geometric evolution of quasi-bilayers in the multicomponent functionalized Cahn-Hilliard equation, J. Math. Biol. 75 (2) (2017) 443–489.
  • [17] K. Promislow, H. Zhang, Critical points of functionalized Lagrangians, Discrete Contin. Dyn. Syst. 33 (4) (2013) 1231–1246.
  • [18] M. Doi, Soft Matter Physics, Oxford University Press, Oxford, UK, 2013.
  • [19] A. Debussche, L. Dettori, On the Cahn-Hilliard equation with a logarithmic free energy, Nonlinear Anal. 24 (10) (1995) 1491–1514.
  • [20] H. Wu, A review on the Cahn-Hilliard equation: classical results and recent advances in dynamic boundary conditions, Electron. Res. Arch. 30 (8) (2022) 2788–2832.
  • [21] H. Abels, M. Wilke, Convergence to equilibrium for the Cahn-Hilliard equation with a logarithmic free energy, Nonlinear Anal. 67 (11) (2007) 3176–3193.
  • [22] C. M. Elliott, H. Garcke, On the Cahn-Hilliard equation with degenerate mobility, SIAM J. Math. Anal. 27 (2) (1996) 404–423.
  • [23] D. Li, A regularization-free approach to the Cahn-Hilliard equation with logarithmic potentials, Discrete Contin. Dyn. Syst. 42 (5) (2022) 2453–2460.
  • [24] A. Miranville, S. Zelik, Robust exponential attractors for Cahn-Hilliard type equations with singular potentials, Math. Methods Appl. Sci. 27 (5) (2004) 545–582.
  • [25] A. Giorgini, M. Grasselli, A. Miranville, The Cahn-Hilliard-Oono equation with singular potential, Math. Models Methods Appl. Sci. 27 (13) (2017) 2485–2510.
  • [26] A. Giorgini, M. Grasselli, H. Wu, The Cahn-Hilliard-Hele-Shaw system with singular potential, Ann. Inst. H. Poincaré Anal. Non Linéaire 35 (4) (2018) 1079–1118.
  • [27] L. Cherfils, A. Miranville, S. Zelik, The Cahn-Hilliard equation with logarithmic potentials, Milan J. Math. 79 (2) (2011) 561–596.
  • [28] W. Chen, J. Jing, C. Wang, X. Wang, S. M. Wise, A modified Crank-Nicolson numerical scheme for the Flory-Huggins Cahn-Hilliard model., Commun. Comput. Phys. 31 (1) (2022) 60–93.
  • [29] W. Chen, C. Wang, X. Wang, S. M. Wise, Positivity-preserving, energy stable numerical schemes for the Cahn-Hilliard equation with logarithmic potential, J. Comput. Phys. X 3 (2019) 100031, 29.
  • [30] M. I. M. Copetti, C. M. Elliott, Numerical analysis of the Cahn-Hilliard equation with a logarithmic free energy, Numer. Math. 63 (1) (1992) 39–65.
  • [31] D. Jeong, J. Kim, A practical numerical scheme for the ternary Cahn-Hilliard system with a logarithmic free energy, Phys. A 442 (2016) 510–522.
  • [32] D. Jeong, S. Lee, J. Kim, An efficient numerical method for evolving microstructures with strong elastic inhomogeneity, Modelling Simul. Mater. Sci. Eng. 23 (4) (2015) Paper No. 045007.
  • [33] H. Jia, Y. Li, G. Feng, K. Li, An efficient two-grid method for the Cahn-Hilliard equation with the concentration-dependent mobility and the logarithmic Flory-Huggins bulk potential, Appl. Math. Comput. 387 (2020) Paper No. 124548, 15 pp.
  • [34] X. Li, Z. Qiao, H. Zhang, An unconditionally energy stable finite difference scheme for a stochastic Cahn-Hilliard equation, Sci. China Math. 59 (9) (2016) 1815–1834.
  • [35] D. Li, T. Tang, Stability of the semi-implicit method for the Cahn-Hilliard equation with logarithmic potentials, Ann. Appl. Math. 37 (1) (2021) 31–60.
  • [36] X. Yang, J. Zhao, On linear and unconditionally energy stable algorithms for variable mobility Cahn-Hilliard type equation with logarithmic Flory-Huggins potential, Commun. Comput. Phys. 25 (3) (2019) 703–728.
  • [37] S. Dai, Q. Liu, K. Promislow, Weak solutions for the functionalized Cahn-Hilliard equation with degenerate mobility, Appl. Anal. 100 (2021) 1–16.
  • [38] S. Dai, Q. Liu, T. Luong, K. Promislow, On nonnegative solutions for the functionalized C-ahn-Hilliard equation with degenerate mobility, Results Appl. Math. 12 (2021) Paper No. 100195, 13 pp.
  • [39] K. Cheng, C. Wang, S. M. Wise, Z. Yuan, Global-in-time Gevrey regularity solutions for the functionalized Cahn-Hilliard equation, Discrete Cont. Dyn. Sys. Ser. S 13 (8) (2020) 2211–2229.
  • [40] A. Miranville, Asymptotic behavior of a sixth-order Cahn–Hilliard system, Cent. Eur. J. Math. 12 (2014) 141–154.
  • [41] N. Duan, Y. Cui, Z. X., A sixth-order phase-field equation with degenerate mobility, Bull. Malays. Math. Sci. Soc. 42 (2019) 79–103.
  • [42] G. Schimperna, H. Wu, On a class of sixth-order Cahn-Hilliard-type equations with logarithmic potential, SIAM J. Math. Anal. 52 (5) (2020) 5155–5195.
  • [43] A. Christlieb, J. Jones, K. Promislow, B. Wetton, M. Willoughby, High accuracy solutions to energy gradient flows from material science models, J. Comput. Phys. 257 (2014) 193–215.
  • [44] R. Guo, Y. Xu, Z. Xu, Local discontinuous Galerkin methods for the functionalized Cahn-Hilliard equation, J. Sci. Comput. 63 (3) (2015) 913–937.
  • [45] W. Feng, Z. Guan, J. Lowengrub, C. Wang, S. M. Wise, Y. Chen, A uniquely solvable, energy stable numerical scheme for the functionalized Cahn-Hilliard equation and its convergence analysis, J. Sci. Comput. 76 (3) (2018) 1938–1967.
  • [46] C. Zhang, J. Ouyang, X. Wang, Y. Chai, M. Ma, Analysis of the energy stability for stabilized semi-implicit schemes of the functionalized Cahn-Hilliard mass-conserving gradient flow equation, J. Sci. Comput. 87 (1) (2021) Paper No. 34, 25.
  • [47] C. Zhang, J. Ouyang, Unconditionally energy stable second-order numerical schemes for the functionalized Cahn-Hilliard gradient flow equation based on the SAV approach, Comput. Math. Appl. 84 (2021) 16–38.
  • [48] C. Zhang, J. Ouyang, X. Wang, S. Li, J. Mao, Highly accurate, linear, and unconditionally energy stable large time-stepping schemes for the functionalized Cahn-Hilliard gradient flow equation, J. Comput. Appl. Math. 392 (2021) Paper No. 113479, 23 pp.
  • [49] C. Zhang, J. Ouyang, C. Wang, S. M. Wise, Numerical comparison of modified-energy stable SAV-type schemes and classical BDF methods on benchmark problems for the functionalized Cahn-Hilliard equation, J. Comput. Phys. 423 (2020) 109772, 35.
  • [50] G. Schimperna, I. Pawłow, On a class of Cahn-Hilliard models with nonlinear diffusion, SIAM J. Math. Anal. 45 (1) (2013) 31–63.
  • [51] C. M. Elliott, A. M. Stuart, The global dynamics of discrete semilinear parabolic equations, SIAM J. Numer. Anal. 30 (6) (1993) 1622–1663.
  • [52] D. J. Eyre, Unconditionally gradient stable time marching the Cahn-Hilliard equation, MRS Online Proceedings Library 529 (1998) 39–46.
  • [53] L. Dong, C. Wang, H. Zhang, Z. Zhang, A positivity-preserving, energy stable and convergent numerical scheme for the Cahn-Hilliard equation with a Flory-Huggins-de Gennes energy, Commun. Math. Sci. 17 (4) (2019) 921–939.
  • [54] L. Dong, C. Wang, H. Zhang, Z. Zhang, A positivity-preserving second-order BDF scheme for the Cahn-Hilliard equation with variable interfacial parameters, Commun. Comput. Phys. 28 (3) (2020) 967–998.
  • [55] C. Liu, C. Wang, S. M. Wise, X. Yue, S. Zhou, A positivity-preserving, energy stable and convergent numerical scheme for the Poisson-Nernst-Planck system, Math. Comput. 90 (331) (2021) 2071–2106.
  • [56] C. Liu, C. Wang, Y. Wang, A structure-preserving, operator splitting scheme for reaction-diffusion equations with detailed balance, J. Comput. Phys. 436 (2021) 110253.
  • [57] J. Simon, Compact sets in the space Lp(0,T;B){L}^{p}(0,{T};{B}), Ann. Mat. Pura Appl. 146 (4) (1987) 65–96.
  • [58] S. M. Wise, Unconditionally stable finite difference, nonlinear multigrid simulation of the Cahn-Hilliard-Hele-Shaw system of equations, J. Sci. Comput. 44 (1) (2010) 38–68.
  • [59] S. M. Wise, C. Wang, J. S. Lowengrub, An energy-stable and convergent finite-difference scheme for the phase field crystal equation, SIAM J. Numer. Anal. 47 (3) (2009) 2269–2288.
  • [60] J. Guo, C. Wang, S. M. Wise, X. Yue, An H2H^{2} convergence of a second-order convex-splitting, finite difference scheme for the three-dimensional Cahn-Hilliard equation, Commun. Math. Sci. 14 (2) (2016) 489–515.
  • [61] C. Wang, S. M. Wise, An energy stable and convergent finite-difference scheme for the modified phase field crystal equation, SIAM J. Numer. Anal. 49 (3) (2011) 945–969.
  • [62] H. O. Kreiss, J. Oliger, Stability of the Fourier method, SIAM J Numer. Anal. 16 (3) (1979) 421–433.
  • [63] L. Wang, W. Chen, C. Wang, An energy-conserving second order numerical scheme for nonlinear hyperbolic equation with an exponential nonlinear term, J. Comput. Appl. Math. 280 (2015) 347–366.
  • [64] X. Li, Z. Qiao, C. Wang, Convergence analysis for a stabilized linear semi-implicit numerical scheme for the nonlocal Cahn–Hilliard equation, Math. Comput. 90 (327) (2021) 171–188.
  • [65] W. Chen, Q. Liu, J. Shen, Error estimates and blow-up analysis of a finite-element approximation for the parabolic-elliptic keller-segel system, arXiv preprint arXiv:2212.07655 (2022).
  • [66] W. Feng, C. Wang, S. M. Wise, Z. Zhang, A second-order energy stable backward differentiation formula method for the epitaxial thin film equation with slope selection, Numer. Methods Partial Differential Eq. 34 (6) (2018) 1975–2007.
  • [67] W. Feng, A. J. Salgado, C. Wang, S. M. Wise, Preconditioned steepest descent methods for some nonlinear elliptic equations involving pp-Laplacian terms, J. Comput. Phys. 334 (2017) 45–67.
  • [68] J. Zhang, C. Wang, S. M. Wise, Z. Zhang, Structure-preserving, energy stable numerical schemes for a liquid thin film coarsening model, SIAM J. Sci. Comput. 43 (2) (2021) A1248–A1272.
  • [69] N. Kraitzman, K. Promislow, An overview of network bifurcations in the functionalized Cahn-Hilliard free energy, in: Mathematics of energy and climate change, Vol. 2 of CIM Ser. Math. Sci., Springer, Cham, 2015, pp. 191–214.
  • [70] S. Jain, F. S. Bates, On the origins of morphological complexity in block copolymer surfactants, Science 300 (5618) (2003) 460–464.