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

Convergence analysis of exponential time differencing scheme for the nonlocal Cahn-Hilliard equation

Danni Zhang zhdanni0807@smail.xtu.edu.cn; Dongling Wang wdymath@xtu.edu.cn; ORCID 0000-0001-8509-2837 School of Mathematics and Computational Science, Xiangtan University, Xiangtan, Hunan 411105, China
Abstract

In this paper, we present a rigorous proof of the convergence of first order and second order exponential time differencing (ETD) schemes for solving the nonlocal Cahn-Hilliard (NCH) equation. The spatial discretization employs the Fourier spectral collocation method, while the time discretization is implemented using ETD-based multistep schemes. The absence of a higher-order diffusion term in the NCH equation poses a significant challenge to its convergence analysis. To tackle this, we introduce new error decomposition formulas and employ the higher-order consistency analysis. These techniques enable us to establish the \ell^{\infty} bound of numerical solutions under some natural constraints. By treating the numerical solution as a perturbation of the exact solution, we derive optimal convergence rates in (0,T;Hh1)2(0,T;2)\ell^{\infty}(0,T;H_{h}^{-1})\cap\ell^{2}(0,T;\ell^{2}). We conduct several numerical experiments to validate the accuracy and efficiency of the proposed schemes, including convergence tests and the observation of long-term coarsening dynamics.

keywords:
Nonlocal Cahn-Hilliard equation, exponential time differencing, Convergence analysis, higher-order consistency analysis.
label1label1footnotetext: The research of Dongling Wang is supported in part by the National Natural Science Foundation of China under grants 12271463 and Outstanding youth fund of department of education of Hunan province under grants 22B0173. The work of Danni Zhang is supported by Hunan Provincial Innovation Foundation For Postgraduate (No: CX20230625).
Declarations of interest: none.

1 Introduction

The Cahn-Hilliard (CH) equation was originally proposed in [6] as one of the most typical phase field models which provides a macroscopic description of phase separation and microstructure evolution in binary alloy systems. As a nonlocal variant of the classic CH equation, the Nonlocal Cahn-Hilliard (NCH) equation has attracted increasing attention and has been widely applied in diverse fields ranging from chemistry, material science and image processing [20, 14, 1, 16, 2]. In this paper, we consider the following NCH equation with the periodic boundary conditions [2, 9]:

ut=Δ(ε2u+f(u)),(x,t)Ω×(0,T],u_{t}=\Delta(\varepsilon^{2}\mathcal{L}u+f(u)),\quad(x,t)\in\Omega\times(0,T], (1.1)

where Ω=i=1d(Xi,Xi)\Omega=\prod_{i=1}^{d}(-X_{i},X_{i}) is a rectangular domain in d(d=1,2,3)\mathbb{R}^{d}(d=1,2,3) and u=u(x,t)u=u(x,t) is an order parameter subject to the initial condition u(x,0)=u0(x)u(x,0)=u_{0}(x), T>0T>0 is the terminal time, ε>0\varepsilon>0 is an interfacial parameter. The function f(u)=F(u)f(u)=F^{\prime}(u) and F(u)=14(u21)2F(u)=\frac{1}{4}(u^{2}-1)^{2} is a double well function. The nonlocal linear operator \mathcal{L} is defined by:

:v(x)ΩJ(xy)(v(x)v(y))dy,\mathcal{L}:v(x)\longmapsto\int_{\Omega}J(x-y)(v(x)-v(y))\,\mathrm{d}y, (1.2)

where JJ is a nonnegative, Ω\Omega-periodic radial kernel function and has a finite second moment in Ω\Omega.

The linear operator \mathcal{L} with the kernel function JJ is self-adjoint and positive semi-definite. Further, if JJ is integrable, then v=(J1)vJv\mathcal{L}v=(J\ast 1)v-J\ast v, where

(Jv)(x)=ΩJ(xy)v(y)dy=ΩJ(y)v(xy)dy.(J*v)(x)=\int_{\Omega}J(x-y)v(y)\,\mathrm{d}y=\int_{\Omega}J(y)v(x-y)\,\mathrm{d}y. (1.3)

We assume γ0:=ϵ2(J1)1>0\gamma_{0}:=\epsilon^{2}(J\ast 1)-1>0 to ensure the NCH equation (1.1) is positive diffusive [4, 3].

The NCH equation (1.1) can be view as the H1H^{-1} gradient flow with respect to the free energy functional

E(u)=ΩF(u(x))dx+ε22(u,u),E(u)=\int_{\Omega}F(u(x))\,\mathrm{d}x+\frac{\varepsilon^{2}}{2}(\mathcal{L}u,u), (1.4)

where (,)(\cdot,\cdot) denotes the standard L2L^{2} inner product on Ω\Omega. For the smooth solutions of (1.1), the total mass is conserved:

ddtΩu(x,t)dx0.\frac{\,\mathrm{d}}{\,\mathrm{d}t}\int_{\Omega}u(x,t)\,\mathrm{d}x\equiv 0. (1.5)

In this paper, we will only consider initial data u0u_{0} with mean zero. Then, the fractional Laplacian operator ||s=(Δ)s2|\nabla|^{s}=(-\Delta)^{\frac{s}{2}} for s<0s<0 is well defined [25]. Due to the gradient structure of (1.1), the energy dissipation laws hold:

ddtE(u)=||1utL2(Ω)20.\frac{\,\mathrm{d}}{\,\mathrm{d}t}E(u)=-\||\nabla|^{-1}u_{t}\|_{L^{2}(\Omega)}^{2}\leq 0. (1.6)

There have been many works on both theoretical analysis and numerical methods for the NCH equation. In mathematical analysis, the well-posedness of the NCH equation equipped with a Neumann or Dirichlet boundary conditions were studied in [3, 4] with an integrable kernel function and in [19] claimed that the existence and uniqueness of the solution to the NCH equation subject to the periodic boundary condition may be established by using a similar technique.

For numerical methods, because of the energetic variational structure inherent in the nonlocal phase field model, an important fact is that the exact solution decrease the energy in time to NCH equation. Therefore, it is essential to develop the so called energy stable numerical algorithms share this key property at the discrete time level. Energy stable numerical schemes of the NCH equation including convex splitting schemes [19, 18], exponential time differencing (ETD) schemes [30, 31, 15], stabilized schemes [9, 27, 28], and so on.

The convergence of numerical solution to NCH equation are established in [19, 17, 26, 27, 28]. In [19, 17], convex splitting schemes were introduced for the NCH equation, demonstrating energy stability and convergence properties. These schemes handled the nonlinear term implicitly and the nonlocal term explicitly, leading to effective numerical solutions but nonlinear iterations were required. Du et al. proposed energy-stable linear semi-implicit schemes in [9], utilizing stabilization techniques to avoid nonlinear iterations. The work in [26] presented the convergence in the discrete H1H^{-1} norm and established the \ell^{\infty} bound of the numerical solutions for the first-order stabilized linear semi-implicit scheme proposed in [9]. Furthermore, Li et al. [27, 28] developed two energy-stable and convergent second-order linear numerical schemes for the NCH equation. These schemes involved combining a modified Crank-Nicolson (CN) approximation with the Adams-Bashforth extrapolation and stability of the second-order backward differentiation formula (BDF2) for the time discretization, resulting in accurate and stable numerical solutions for the NCH equation.

However, the convergence analysis of ETD schemes for the NCH equation is more challenging and there are currently no results, one of the reasons being that the lack of higher-order diffusion term and the Laplacian of nonlinear term. The ETD schemes [5, 7, 12, 13, 21], which involve exact integration of the target equation followed by an explicit approximation of the temporal integral of the nonlinear term. Exact evaluation of the linear terms makes the ETD schemes achieve high accuracy and satisfactory stability when dealing with stiff differential equations. The first- and second-order ETD schemes were applied to the Nonlocal Allen-Cahn equation and have been analyzed energy stable and \ell^{\infty} convergence [10], and further extended to a class of semi-linear parabolic equations [11]. Convergence and the \ell^{\infty} bound of numerical solutions were proved of ETD schemes to classic CH equation [24].

Compare with the classic CH equation or Nonlocal Allen-Cahn equation, the NCH equation equipped with nonlocal diffusion operator and has more complicated nonlinear term, so that the analytical techniques mentioned above are hardly applicable to the NCH equation. In this paper, we establish the convergence of the fully discrete first-order ETD scheme (ETD1) and second-order ETD multistep scheme (ETD2) for the NCH equation which are developed our recent work [31]. Our analysis is mainly based on two important observations:

  • 1.

    Unlike the standard error analysis based on the error equation (3.5), our analysis based on two new error decomposition formulas

    en:=e~n+τ𝒫Nuτ,1+τ2𝒫Nuτ,2(ETD1 scheme);\displaystyle e^{n}:=\tilde{e}^{n}+\tau\mathcal{P}_{N}u_{\tau,1}+\tau^{2}\mathcal{P}_{N}u_{\tau,2}\quad(\text{ETD1 scheme});
    en:=e~n+τ2𝒫Nuτ,3(ETD2 scheme).\displaystyle e^{n}:=\tilde{e}^{n}+\tau^{2}\mathcal{P}_{N}u_{\tau,3}\hskip 52.63777pt(\text{ETD2 scheme}).

    See the details in (3.9). This formula allows us to establish the higher order consistency and obtain the \ell^{\infty} bound of the numerical solutions.

  • 2.

    We adopt (ΔN)1e~n+1(-\Delta_{N})^{-1}\tilde{e}^{n+1} to test the error functions in (3.23) and (3.2.2), instead of e~n+1\tilde{e}^{n+1} as in a previous work [24], where (ΔN)1(-\Delta_{N})^{-1} is a spatial discrete operator defined in the Appendix A. This method can be used to overcome the possible instability caused by the absence of high-order diffusion terms in the NCH equation.

Based on the above two points, we prove the convergence in the discrete H1H^{-1} norm and obtain the \ell^{\infty} bound of numerical solutions under some natural constraints on the space-time step sizes.

The rest of this paper is organized as follows. The fully discrete ETD1 and ETD2 schemes are constructed in Section 2, and some notations, definitions and useful lemmas are also introduced. In Section 3, we prove the convergence by the higher-order consistency estimate. In addition, the \ell^{\infty} bound of numerical solutions are also obtained. In Section 4, some numerical experiments are carried out to test the convergence in time level. The coarsening dynamics are simulated to show the long time behaviors of ETD2. Some concluding remarks are given in Section 5.

2 Fully discrete exponential time differencing schemes

In this section, we present the fully discrete exponential time differencing schemes for the NCH equation, where the Fourier spectral collocation method is adopted for the spatial discretization.

We consider the square domain Ω=(X,X)2\Omega=(-X,X)^{2} and define the spatial grid Ωh={(xi,yj)=(X+ih,X+jh), 1i,jN},\Omega_{h}=\{(x_{i},y_{j})=(-X+ih,-X+jh),\;1\leq i,j\leq N\}, where the space step size h=2XNh=\frac{2X}{N} (NN is even). We need two spaces. Let h\mathcal{M}_{h} be the space of all the Ωh\Omega_{h}-periodic grid functions. That means if fhf\in\mathcal{M}_{h}, then ff is a discrete function defined on the discrete grid Ωh\Omega_{h}. Let h0\mathcal{M}_{h}^{0} be the space of the zero-mean grid functions. That means if fh0f\in\mathcal{M}_{h}^{0}, then f¯:=h24X2(i,j)Shfi,j=0\overline{f}:=\frac{h^{2}}{4X^{2}}\sum_{(i,j)\in S_{h}}f_{i,j}=0, where ShS_{h} is the index set defined in the Appendix A.

We denote ΔN\Delta_{N} and N\mathcal{L}_{N} are Fourier spectral collocation space approximation operators of Δ\Delta and \mathcal{L}. These detailed definitions and some related properties are given in the Appendix A. For any f,ghf,g\in\mathcal{M}_{h}, the discrete inner product ,\langle\cdot,\cdot\rangle, ,1,N\langle\cdot,\cdot\rangle_{-1,N} and norm 2\|\cdot\|_{2}, \|\cdot\|_{\infty}, 1,N\|\cdot\|_{-1,N}, and the discrete convolution fϕf\circledast\phi are defined in the Appendix A.

It is well known [12] that a suitable linear operator splitting can improve the stability. So we can introduce a stabilizing parameter κ>0\kappa>0 and define

Lh:=ε2ΔNNκΔN,fκ(U):=f(U)κU.L_{h}:=-\varepsilon^{2}\Delta_{N}\mathcal{L}_{N}-\kappa\Delta_{N},\;\,\;f_{\kappa}(U):=f(U)-\kappa U. (2.1)

Now LhL_{h} is self-adjoint and positive definite on h0\mathcal{M}_{h}^{0}. Then the space semi-discrete equation for (1.1) is to find U:[0,T]h0U:[0,T]\to\mathcal{M}_{h}^{0} such that

{dUdt+LhU=ΔNfκ(U),t(0,T],U(0)=U0,\begin{split}\left\{\begin{split}&\frac{\,\mathrm{d}U}{\,\mathrm{d}t}+L_{h}U=\Delta_{N}f_{\kappa}(U),\quad t\in(0,T],\\ &U(0)=U^{0},\end{split}\right.\end{split} (2.2)

where the initial value U0h0U^{0}\in\mathcal{M}_{h}^{0} is given.

Let {tn=nτ:0nNt}\{t_{n}=n\tau:0\leq n\leq N_{t}\}, where τ=TNt\tau=\frac{T}{N_{t}} is the time step size and NtN_{t} is a given positive integer. By using the property of the differentiation of matrix exponentials, the solution of the equation (2.2) satisfies

U(tn+1)=eτLhU(tn)+0τe(τs)LhΔNfκ(U(tn+s))ds.U(t_{n+1})=e^{-\tau L_{h}}U(t_{n})+\int_{0}^{\tau}e^{-(\tau-s)L_{h}}\Delta_{N}f_{\kappa}(U(t_{n}+s))\,\mathrm{d}s. (2.3)

The key to construct the ETD schemes is to approximate the nonlinear term fκ(U(tn+s))f_{\kappa}(U(t_{n}+s)) in (2.3) by polynomial interpolation, and then precisely integrate the interpolation.

The ETD1 scheme comes from approximating fκ(U(tn+s))f_{\kappa}(U(t_{n}+s)) by the constant fκ(U(tn))f_{\kappa}(U(t_{n})), given by

Un+1=ϕ1(τLh)Un+τϕ0(τLh)ΔNfκ(Un),U^{n+1}=\phi_{-1}(\tau L_{h})U^{n}+\tau\phi_{0}(\tau L_{h})\Delta_{N}f_{\kappa}(U^{n}), (2.4)

where ϕ1(a)=ea,ϕ0(a)=1eaa,a0\phi_{-1}(a)=e^{-a},\phi_{0}(a)=\frac{1-e^{-a}}{a},a\neq 0. The ETD2 scheme is obtained by approximating fκ(U(tn+s))f_{\kappa}(U(t_{n}+s)) by a linear extrapolation based on fκ(U(tn))f_{\kappa}(U(t_{n})) and fκ(U(tn1))f_{\kappa}(U(t_{n-1})), given by

Un+1=ϕ1(τLh)Un+τ[(ϕ0+ϕ1)(τLh)ΔNfκ(Un)ϕ1(τLh)ΔNfκ(Un1)],U^{n+1}=\phi_{-1}(\tau L_{h})U^{n}+\tau[(\phi_{0}+\phi_{1})(\tau L_{h})\Delta_{N}f_{\kappa}(U^{n})-\phi_{1}(\tau L_{h})\Delta_{N}f_{\kappa}(U^{n-1})], (2.5)

where ϕ1(a)=a1+eaa2,a0.\phi_{1}(a)=\frac{a-1+e^{-a}}{a^{2}},a\neq 0.

The ETD2 scheme (2.5) is a two-step algorithm, an accurate approximation for the value at t1t_{1} is needed. Usually, U1U^{1} can be computed by the ETD1. But in this paper, we choose a higher-order approximation for U1U^{1}, to facilitate the higher-order consistency analysis presented in the later sections. For example, we can apply the second-order Runge-Kutta (RK2) method in first step, which yields a third-order temporal accuracy at t1t_{1} for exact initial data U0U^{0}.

The following lemmas are useful in our following convergence analysis.

Lemma 2.1 ([24]).

(i) For a>0a>0, the following inequalities hold: 0<(1+a)ϕ1(a)<1, 1<(1+a)ϕ0(a)<32,12<(1+a)ϕ1(a)<10<(1+a)\phi_{-1}(a)<1,\,1<(1+a)\phi_{0}(a)<\frac{3}{2},\,\frac{1}{2}<(1+a)\phi_{1}(a)<1 and 0<(1+a)[ϕ0(a)ϕ1(a)]<1.0<(1+a)[\phi_{0}(a)-\phi_{1}(a)]<1.

(ii) If 0<s<τ10<s<\tau\leq 1, then for any a>0a>0, it holds 0<(1+aτ)ea(τs)<1.0<(1+a\tau)e^{-a(\tau-s)}<1.

Lemma 2.2 ([9]).

The operator N\mathcal{L}_{N} has the following properties:

(i) the eigenvalues of N\mathcal{L}_{N} are λkl=h2(J^00J^kl)0,(k,l)S^h\lambda_{kl}=h^{2}(\hat{J}_{00}-\hat{J}_{kl})\geq 0,\,(k,l)\in\widehat{S}_{h};

(ii) N\mathcal{L}_{N} commutes with ΔN\Delta_{N} and is self-adjoint and positive semi-define;

(iii) for any fhf\in\mathcal{M}_{h}, we have Nf=(J1)fJf\mathcal{L}_{N}f=(J\circledast 1)f-J\circledast f.

Lemma 2.3 ([26]).

Suppose JCper1(Ω)J\in C_{per}^{1}(\Omega) and define its gird restriction via Jij:=J(xi,yj)J_{ij}:=J(x_{i},y_{j}). Then for any ϕ,ψh\phi,\psi\in\mathcal{M}_{h} and α>0\alpha>0, we have

|Jϕ,ΔNψ|αϕ22+CJαNψ22,|\langle J\circledast\phi,\Delta_{N}\psi\rangle|\leq\alpha\|\phi\|_{2}^{2}+\frac{C_{J}}{\alpha}\|\nabla_{N}\psi\|_{2}^{2}, (2.6)

where CJC_{J} is a positive constant that depends on JJ and Ω\Omega but is independent of hh.

3 Convergence analysis

In this section, we analyze the convergence of the ETD1 scheme and ETD2 scheme, and give an optimal rate error estimates in (0,T;Hh1)2(0,T,2)\ell^{\infty}(0,T;H_{h}^{-1})\cap\ell^{2}(0,T,\ell^{2}). To perform the convergence analysis, let’s first introduce the concept and preliminary.

For a linear symmetric operator Q:hhQ:\mathcal{M}_{h}\to\mathcal{M}_{h}, we define the norm Q\interleave Q\interleave by the spectrum radius of QQ, i.e., Q=max{|λ|:λσ(Q)}\interleave Q\interleave=\max\{|\lambda|:\lambda\in\sigma(Q)\}, where σ(Q)\sigma(Q) is the set of all the eigenvalues of QQ. It holds that Qv2Qv2,vh.\|Qv\|_{2}\leq\interleave Q\interleave\cdot\|v\|_{2},\;\forall v\in\mathcal{M}_{h}.

Denote by ueu_{e} the exact solution of (1.1). The existence and uniqueness of ueu_{e} may be obtained by using the techniques adopted in [3, 4], from which one can get the following estimate

ueL(0,T;L(Ω))+ueL(0,T;L(Ω))C,T>0.\|u_{e}\|_{L^{\infty}(0,T;L^{\infty}(\Omega))}+\|\nabla u_{e}\|_{L^{\infty}(0,T;L^{\infty}(\Omega))}\leq C,\quad\forall T>0. (3.1)

Moreover, if the initial data is sufficient regularity, we can assume that the exact solution has regularity as

ue:=H4(0,T;Cper0(Ω¯))H3(0,T;Cper2(Ω¯))L(0,T;Cperm+2(Ω¯)),m3.u_{e}\in\mathcal{R}:=H^{4}\left(0,T;C^{0}_{per}(\bar{\Omega})\right)\cap H^{3}\left(0,T;C^{2}_{per}(\bar{\Omega})\right)\cap L^{\infty}\left(0,T;C^{m+2}_{per}(\bar{\Omega})\right),\quad m\geq 3. (3.2)

Let K\mathcal{B}^{K} be the space of trigonometric polynomials of degree up to K:=N2K:=\frac{N}{2}. Let 𝒫N:L2(Ω)K\mathcal{P}_{N}:L^{2}(\Omega)\to\mathcal{B}^{K} be the L2L^{2} orthogonal projection operator. We define uN(,t):=𝒫Nue(,t)u_{N}(\cdot,t):=\mathcal{P}_{N}u_{e}(\cdot,t). If ueL(0,T;Hper(Ω))u_{e}\in L^{\infty}(0,T;H^{\ell}_{per}(\Omega)) for some \ell\in\mathbb{N}, the following estimate is standard [29]

uNueL(0,T;Hk(Ω))ChkueL(0,T;H(Ω)),0k.\|u_{N}-u_{e}\|_{L^{\infty}(0,T;H^{k}(\Omega))}\leq Ch^{\ell-k}\|u_{e}\|_{L^{\infty}(0,T;H^{\ell}(\Omega))},\;\;\;\forall 0\leq k\leq\ell. (3.3)

By the orthogonality of Fourier projection 𝒫N\mathcal{P}_{N}, we have ΩuN(,tn)dx=Ωue(,tn)dx\int_{\Omega}u_{N}(\cdot,t_{n})\,\mathrm{d}x=\int_{\Omega}u_{e}(\cdot,t_{n})\,\mathrm{d}x. Noting that the exact solution ueu_{e} is mass conservative, i.e., Ωue(,tn)dx=Ωue(,tn1)dx\int_{\Omega}u_{e}(\cdot,t_{n})\,\mathrm{d}x=\int_{\Omega}u_{e}(\cdot,t_{n-1})\,\mathrm{d}x. Thus, we obtain

ΩuN(,tn)dx=ΩuN(,tn1)dx,n.\int_{\Omega}u_{N}(\cdot,t_{n})\,\mathrm{d}x=\int_{\Omega}u_{N}(\cdot,t_{n-1})\,\mathrm{d}x,\quad\forall n\in\mathbb{N}. (3.4)

Since uNKu_{N}\in\mathcal{B}^{K}, we have ΩuN(,tn)dx=h2(i,j)ShuN(xi,yj,tn)\int_{\Omega}u_{N}(\cdot,t_{n})\,\mathrm{d}x=h^{2}\sum_{(i,j)\in S_{h}}u_{N}(x_{i},y_{j},t_{n}), that means the rectangular quadrature rule holds exactly for all uNKu_{N}\in\mathcal{B}^{K}. The values of uN(tn)u_{N}(t_{n}) at discrete grid points are still denoted as uN(tn)u_{N}(t_{n}), i.e., uN(xi,yj,tn)=𝒫Nue(tn)|i,ju_{N}(x_{i},y_{j},t_{n})=\mathcal{P}_{N}u_{e}(t_{n})|_{i,j}. Recall  uN¯(tn)=h24X2(i,j)ShuN(xi,yj,tn),\overline{u_{N}}(t_{n})=\frac{h^{2}}{4X^{2}}\sum_{(i,j)\in S_{h}}u_{N}(x_{i},y_{j},t_{n}), then from (3.4) we can get the mass conservative property of uNu_{N} in the discrete average sense: uN¯(tn)=uN¯(tn1).\overline{u_{N}}(t_{n})=\overline{u_{N}}(t_{n-1}).

On the other hand, the solutions of the numerical schemes (2.4)-(2.5) are also mass conservative [30]: Un¯=Un1¯,n.\overline{U^{n}}=\overline{U^{n-1}},\;\forall n\in\mathbb{N}. We use the mass conservative for the initial value U0=uN(t0)U^{0}=u_{N}(t_{0}), and define the error grid function

en:=UnuN(tn),n0.e^{n}:=U^{n}-u_{N}(t_{n}),\quad\forall n\geq 0. (3.5)

We have en¯=0\overline{e^{n}}=0, so en1,N\|e^{n}\|_{-1,N} is well defined. Under the regularity assumption (3.1), for the projection functions uN(tk):=𝒫Nue(tk)u_{N}(t_{k}):=\mathcal{P}_{N}u_{e}(t_{k}), we have

max1kNtuN(tk)+max1kNtNuN(tk)C.\max_{1\leq k\leq N_{t}}\|u_{N}(t_{k})\|_{\infty}+\max_{1\leq k\leq N_{t}}\|\nabla_{N}u_{N}(t_{k})\|_{\infty}\leq C. (3.6)

The main result of this work is the following theorem.

Theorem 3.1.

Assume that the solution of the NCH equation (1.1) satisfying the regularity class \mathcal{R} presented in (3.2). Also assume that τ\tau and hh are small sufficiently and satisfy τCh\tau\leq Ch for some constant C>0C>0. Let B:=max1kNtuN(tk)+1B:=\max\limits_{1\leq k\leq N_{t}}\|u_{N}(t_{k})\|_{\infty}+1.

(i) For ETD1 scheme, if the stabilizing parameter κ2B21\kappa\geq 2B^{2}-1, we have

en1,N+(γ1τk=1nek22)1/2C(τ+hm),0nNt;\|e^{n}\|_{-1,N}+\left(\gamma_{1}\tau\sum_{k=1}^{n}\|e^{k}\|_{2}^{2}\right)^{1/2}\leq C(\tau+h^{m}),\quad 0\leq n\leq N_{t}; (3.7)

(ii) For ETD2 scheme, if the stabilizing parameter κ3B21\kappa\geq 3B^{2}-1, we have

en1,N+(γ2τk=1nek22)1/2C(τ2+hm),0nNt,\|e^{n}\|_{-1,N}+\left(\gamma_{2}\tau\sum_{k=1}^{n}\|e^{k}\|_{2}^{2}\right)^{1/2}\leq C(\tau^{2}+h^{m}),\quad 0\leq n\leq N_{t}, (3.8)

where C,γ1,γ2C,\gamma_{1},\gamma_{2} are some positive constant and are independent of τ\tau and hh.

We provide a heuristic explanation of the main idea behind the proof. The usual convergence analysis starts directly from the error equation en:=UnuN(tn)e^{n}:=U^{n}-u_{N}(t_{n}) defined in (3.5). However, our here adopts a new strategy. We construct two new error decomposition representation formulas

en:=e~n+τ𝒫Nuτ,1+τ2𝒫Nuτ,2(ETD1 scheme);en:=e~n+τ2𝒫Nuτ,3(ETD2 scheme),\begin{split}&e^{n}:=\tilde{e}^{n}+\tau\mathcal{P}_{N}u_{\tau,1}+\tau^{2}\mathcal{P}_{N}u_{\tau,2}\quad(\text{ETD1 scheme});\\ &e^{n}:=\tilde{e}^{n}+\tau^{2}\mathcal{P}_{N}u_{\tau,3}\hskip 52.63777pt(\text{ETD2 scheme}),\end{split} (3.9)

where e~n:=Unu~(tn)\tilde{e}^{n}:=U^{n}-\tilde{u}(t_{n}) and u~(tn)\tilde{u}(t_{n}) is the constructed approximate solution defined in (3.14) for the ETD1 scheme and in (3.37) for the ETD2 scheme, uτ,1u_{\tau,1}, uτ,2u_{\tau,2} and uτ,3u_{\tau,3} are temporal correction functions. By doing so, we can obtain an 𝒪(τ3+hm)\mathcal{O}(\tau^{3}+h^{m}) truncation error. This higher consistency allows us to derive a higher order convergence estimate of the modified error function e~n+1\tilde{e}^{n+1} as e~n+11,N=u~(tn+1)Un+11,N=𝒪(τ3+hm)\|\tilde{e}^{n+1}\|_{-1,N}=\|\tilde{u}(t_{n+1})-U^{n+1}\|_{-1,N}=\mathcal{O}(\tau^{3}+h^{m}) (see (3.31) and (3.49)), which in turn lead to Un+1<\|U^{n+1}\|_{\infty}<\infty under reasonable condition τCh\tau\leq Ch. Due to the lack of higher-order diffusion term, we adopt (ΔN)1e~n+1(-\Delta_{N})^{-1}\tilde{e}^{n+1} to test the error function in (3.23) and (3.2.2), instead of e~n+1\tilde{e}^{n+1} as that in [24]. Then, we estimate both sides of the error equation by using the result of the higher-order consistency analysis and induction argument, we can obtain an optimal convergence rate in (0,T;Hh1)2(0,T,2)\ell^{\infty}(0,T;H_{h}^{-1})\cap\ell^{2}(0,T,\ell^{2}) for en+1{e}^{n+1} of the proposed two numerical schemes.

3.1 Convergence analysis of ETD1

3.1.1 Higher-order consistency analysis

For a given UnU^{n}, the solution Un+1U^{n+1} of the ETD1 scheme (2.4) can be given by Un+1=W(τ)U^{n+1}=W(\tau) with the function W:[0,τ]h0W:[0,\tau]\to\mathcal{M}_{h}^{0} determined by the following evolution equation:

{dW(s)ds+LhW(s)=ΔNfκ(Un),s(0,τ),W(0)=Un.\begin{split}\left\{\begin{split}&\frac{\,\mathrm{d}W(s)}{\,\mathrm{d}s}+L_{h}W(s)=\Delta_{N}f_{\kappa}(U^{n}),\quad s\in(0,\tau),\\ &W(0)=U^{n}.\end{split}\right.\end{split} (3.10)

Then, for the continuous NCH equation (1.1), we can give a similar expression as follows. For given ue(x,tn)u_{e}(x,t_{n}), the solution ue(x,tn+1)u_{e}(x,t_{n+1}) can be obtained by ue(x,tn+1)=w(x,τ)u_{e}(x,t_{n+1})=w(x,\tau) with the function w(x,s)w(x,s) satisfying

{w(s)s=ε2Δw+κΔw+Δfκ(w),xΩ,s(0,τ),w(x,0)=ue(x,tn).\begin{split}\left\{\begin{split}&\frac{\partial w(s)}{\partial s}=\varepsilon^{2}\Delta\mathcal{L}w+\kappa\Delta w+\Delta f_{\kappa}(w),\quad x\in\Omega,s\in(0,\tau),\\ &w(x,0)=u_{e}(x,t_{n}).\end{split}\right.\end{split} (3.11)

Denote the function WN(s)=𝒫Nw(s)W_{N}(s)=\mathcal{P}_{N}w(s). Then WN(s)W_{N}(s) satisfies the following equation:

WN(s)s=ε2ΔWN(s)+κΔWN(s)+Δ𝒫N(ue(tn+s)3)ΔWN(s)κΔWN(s),\frac{\partial W_{N}(s)}{\partial s}=\varepsilon^{2}\Delta\mathcal{L}W_{N}(s)+\kappa\Delta W_{N}(s)+\Delta\mathcal{P}_{N}(u_{e}(t_{n}+s)^{3})-\Delta W_{N}(s)-\kappa\Delta W_{N}(s), (3.12)

and WN(s)W_{N}(s) also solves the discrete equation (3.10) with the local truncation error Rhτ(1)(s)R_{h\tau}^{(1)}(s)

{dWN(s)ds+LhWN(s)=ΔNfκ(uN(tn))+Rhτ(1)(s),s(0,τ),WN(0)=uN(tn).\begin{split}\left\{\begin{split}&\frac{\,\mathrm{d}W_{N}(s)}{\,\mathrm{d}s}+L_{h}W_{N}(s)=\Delta_{N}f_{\kappa}(u_{N}(t_{n}))+R_{h\tau}^{(1)}(s),\quad s\in(0,\tau),\\ &W_{N}(0)=u_{N}(t_{n}).\end{split}\right.\end{split} (3.13)

By comparing equations (3.12) and (3.13), we can get Rhτ(1)(s)=Rh(1)+Rτ(1),R_{h\tau}^{(1)}(s)=R_{h}^{(1)}+R_{\tau}^{(1)}, where

Rh(1)=(ε2Δε2ΔNN)uN(tn+s)+(κΔκΔN)uN(tn+s)\displaystyle R_{h}^{(1)}=(\varepsilon^{2}\Delta\mathcal{L}-\varepsilon^{2}\Delta_{N}\mathcal{L}_{N})u_{N}(t_{n}+s)+(\kappa\Delta-\kappa\Delta_{N})u_{N}(t_{n}+s)
+Δ(𝒫N(ue(tn+s)3)uN(tn+s)3)+Δfκ(uN(tn+s))ΔNfκ(uN(tn+s)),\displaystyle\hskip 28.45274pt+\Delta(\mathcal{P}_{N}(u_{e}(t_{n}+s)^{3})-u_{N}(t_{n}+s)^{3})+\Delta f_{\kappa}(u_{N}(t_{n}+s))-\Delta_{N}f_{\kappa}(u_{N}(t_{n}+s)),
Rτ(1)=ΔNfκ(uN(tn+s))ΔNfκ(uN(tn)).\displaystyle R_{\tau}^{(1)}=\Delta_{N}f_{\kappa}(u_{N}(t_{n}+s))-\Delta_{N}f_{\kappa}(u_{N}(t_{n})).

By the standard Fourier spectral approximation, Rh(1)R_{h}^{(1)} has the bound ChmCh^{m}. By Taylor expansion with the integral remainder, Rτ(1)R_{\tau}^{(1)} can be bounded by CτC\tau. Thus, we have

sups(0,τ)Rhτ(1)(s)1,NC(τ+hm).\sup\limits_{s\in(0,\tau)}\|R_{h\tau}^{(1)}(s)\|_{-1,N}\leq C(\tau+h^{m}).

Since the local truncation error only has first order accuracy in time, it is not enough to derive the boundedness of numerical solution, i.e., Un+1<\|U^{n+1}\|_{\infty}<\infty. To remedy this, we introduce the auxiliary profile

u~=uN+τ𝒫Nuτ,1+τ2𝒫Nuτ,2,\tilde{u}=u_{N}+\tau\mathcal{P}_{N}u_{\tau,1}+\tau^{2}\mathcal{P}_{N}u_{\tau,2}, (3.14)

where uτ,1u_{\tau,1} and uτ,2u_{\tau,2} are two temporal correction functions will be constructed later. By doing so, we can obtain a higher 𝒪(τ3+hm)\mathcal{O}(\tau^{3}+h^{m}) consistency.

Let’s construct the correction function uτ,1u_{\tau,1} first. By Fourier projection from continuous NCH equation (1.1) and carry out the space discretization, we have uN:(tn,tn+1]h0u_{N}:(t_{n},t_{n+1}]\to\mathcal{M}_{h}^{0} satisfying

duNdt+LhuN=ΔNfκ(uN(t))+𝒪(hm),t(tn,tn+1].\frac{\,\mathrm{d}u_{N}}{\,\mathrm{d}t}+L_{h}u_{N}=\Delta_{N}f_{\kappa}(u_{N}(t))+\mathcal{O}(h^{m}),\quad t\in(t_{n},t_{n+1}].

Expansing the nonlinear term fκ(uN(t))f_{\kappa}(u_{N}(t)) for t(tn,tn+1]t\in(t_{n},t_{n+1}] at tnt_{n} to obtain fκ(uN(t))=fκ(uN(tn))+τg1+𝒪(τ2)f_{\kappa}(u_{N}(t))=f_{\kappa}(u_{N}(t_{n}))+\tau g_{1}+\mathcal{O}(\tau^{2}), where g1g_{1} depends only on fκf_{\kappa} and uNu_{N}. We get that

duNdt+LhuN=ΔNfκ(uN(tn))+τg1+𝒪(τ2)+𝒪(hm),t(tn,tn+1].\frac{\,\mathrm{d}u_{N}}{\,\mathrm{d}t}+L_{h}u_{N}=\Delta_{N}f_{\kappa}(u_{N}(t_{n}))+\tau g_{1}+\mathcal{O}(\tau^{2})+\mathcal{O}(h^{m}),\quad t\in(t_{n},t_{n+1}]. (3.15)

By mass conservative and the periodic boundary condition, we also have g1¯=0\overline{g_{1}}=0.

The first order temporal correction function uτ,1u_{\tau,1} is given by solving the following ODE system:

{duτ,1dt+Lhuτ,1=ΔN[fκ(uN(tn))uτ,1(tn)]g1,t(tn,tn+1],uτ,1(0)0,\begin{split}\left\{\begin{split}&\frac{\,\mathrm{d}u_{\tau,1}}{\,\mathrm{d}t}+L_{h}u_{\tau,1}=\Delta_{N}\left[f^{\prime}_{\kappa}(u_{N}(t_{n}))u_{\tau,1}(t_{n})\right]-g_{1},\quad t\in(t_{n},t_{n+1}],\\ &u_{\tau,1}(0)\equiv 0,\end{split}\right.\end{split} (3.16)

The existence and uniqueness of the solution of this linear system is standard and it depends only on uNu_{N}.

Let u~1=uN+τ𝒫Nuτ,1\tilde{u}_{1}=u_{N}+\tau\mathcal{P}_{N}u_{\tau,1}. Then it follows from the equations (3.15) and (3.16) that

du~1dt+Lhu~1=ΔNfκ(u~1(tn))+τ2g2+𝒪(τ3+hm),t(tn,tn+1],\frac{\,\mathrm{d}\tilde{u}_{1}}{\,\mathrm{d}t}+L_{h}\tilde{u}_{1}=\Delta_{N}f_{\kappa}(\tilde{u}_{1}(t_{n}))+\tau^{2}g_{2}+\mathcal{O}(\tau^{3}+h^{m}),\quad t\in(t_{n},t_{n+1}], (3.17)

where we have used the following expansion

fκ(u~1(tn))\displaystyle f_{\kappa}(\tilde{u}_{1}(t_{n})) =fκ(uN(tn))+τfκ(uN(tn))𝒫Nuτ,1(tn)+𝒪(τ2)\displaystyle=f_{\kappa}(u_{N}(t_{n}))+\tau f_{\kappa}^{\prime}(u_{N}(t_{n}))\mathcal{P}_{N}u_{\tau,1}(t_{n})+\mathcal{O}(\tau^{2})
=fκ(uN(tn))+τ𝒫N(fκ(uN(tn))𝒫Nuτ,1(tn))+𝒪(τ2+hm),\displaystyle=f_{\kappa}(u_{N}(t_{n}))+\tau\mathcal{P}_{N}(f_{\kappa}^{\prime}(u_{N}(t_{n}))\mathcal{P}_{N}u_{\tau,1}(t_{n}))+\mathcal{O}(\tau^{2}+h^{m}),

and g2g_{2} defined in (3.17) depends only on fκf_{\kappa} and uNu_{N}. We also have g2¯=0\overline{g_{2}}=0. Similarly, we can get the second order temporal correction function uτ,2u_{\tau,2}, which is the solution of the following linear ODE system:

{duτ,2dt+Lhuτ,2=ΔN[fκ(uN(tn))uτ,2(tn)]g2,t(tn,tn+1],uτ,2(0)0,\begin{split}\left\{\begin{split}&\frac{\,\mathrm{d}u_{\tau,2}}{\,\mathrm{d}t}+L_{h}u_{\tau,2}=\Delta_{N}[f_{\kappa}^{\prime}(u_{N}(t_{n}))u_{\tau,2}(t_{n})]-g_{2},\quad t\in(t_{n},t_{n+1}],\\ &u_{\tau,2}(0)\equiv 0,\end{split}\right.\end{split} (3.18)

The unique solution only dependent on uNu_{N}. Define a grid function u~=u~1+τ2𝒫Nuτ,2\tilde{u}=\tilde{u}_{1}+\tau^{2}\mathcal{P}_{N}u_{\tau,2}. Combination of (3.17) and a Fourier projection of (3.18) leads to

du~dt+Lhu~=ΔNfκ(u~(tn))+𝒪(τ3+hm),t(tn,tn+1],\frac{\,\mathrm{d}\tilde{u}}{\,\mathrm{d}t}+L_{h}\tilde{u}=\Delta_{N}f_{\kappa}(\tilde{u}(t_{n}))+\mathcal{O}(\tau^{3}+h^{m}),\quad t\in(t_{n},t_{n+1}], (3.19)

where we have used the fact

fκ(u~(tn))\displaystyle f_{\kappa}(\tilde{u}(t_{n})) =fκ(u~1(tn))+τ2fκ(uN(tn))𝒫Nuτ,2(tn)+𝒪(τ3)\displaystyle=f_{\kappa}(\tilde{u}_{1}(t_{n}))+\tau^{2}f_{\kappa}^{\prime}(u_{N}(t_{n}))\mathcal{P}_{N}u_{\tau,2}(t_{n})+\mathcal{O}(\tau^{3})
=fκ(u~1(tn))+τ2𝒫N(fκ(uN(tn))𝒫Nuτ,2(tn))+𝒪(τ3+hm).\displaystyle=f_{\kappa}(\tilde{u}_{1}(t_{n}))+\tau^{2}\mathcal{P}_{N}(f_{\kappa}^{\prime}(u_{N}(t_{n}))\mathcal{P}_{N}u_{\tau,2}(t_{n}))+\mathcal{O}(\tau^{3}+h^{m}).

We see that u~¯=0\overline{\tilde{u}}=0. Thus, we have completed the construction of u~\tilde{u}.

Note that u~\|\tilde{u}\|_{\infty} is also bounded. In fact, it follows from uτ,1C\|u_{\tau,1}\|_{\infty}\leq C and uτ,2C\|u_{\tau,2}\|_{\infty}\leq C that u~uNCτ.\|\tilde{u}-u_{N}\|_{\infty}\leq C\tau. In particular, when τ\tau is sufficiently small we have u~uNCτ12\|\tilde{u}-u_{N}\|_{\infty}\leq C\tau\leq\frac{1}{2}, which gives u~uN+u~uNuN+12B.\|\tilde{u}\|_{\infty}\leq\|u_{N}\|_{\infty}+\|\tilde{u}-u_{N}\|_{\infty}\leq\|u_{N}\|_{\infty}+\frac{1}{2}\leq B.

3.1.2 Convergence analysis in (0,T;Hh1)2(0,T;2)\ell^{\infty}\left(0,T;H_{h}^{-1}\right)\cap\ell^{2}\left(0,T;\ell^{2}\right)

Let W~(s)=u~(tn+s)\tilde{W}(s)=\tilde{u}(t_{n}+s) for s[0,τ]s\in[0,\tau]. Then, by consistency, W~(s)\tilde{W}(s) satisfies

{dW~(s)ds+LhW~(s)=ΔNfκ(u~(tn))+R~hτ(1)(s),s(0,τ),W~(0)=u~(tn),\begin{split}\left\{\begin{split}&\frac{\,\mathrm{d}\tilde{W}(s)}{\,\mathrm{d}s}+L_{h}\tilde{W}(s)=\Delta_{N}f_{\kappa}(\tilde{u}(t_{n}))+\tilde{R}_{h\tau}^{(1)}(s),\quad s\in(0,\tau),\\ &\tilde{W}(0)=\tilde{u}(t_{n}),\end{split}\right.\end{split} (3.20)

where the truncation error R~hτ(1)(s)\tilde{R}_{h\tau}^{(1)}(s) satisfies

sups(0,τ)R~hτ(1)(s)1,NC(τ3+hm),\sup\limits_{s\in(0,\tau)}\|\tilde{R}_{h\tau}^{(1)}(s)\|_{-1,N}\leq C(\tau^{3}+h^{m}), (3.21)

Subtracting (3.20) from (3.10) and putting e~(s):=W(s)W~(s)\tilde{e}(s):=W(s)-\tilde{W}(s) yield

{de~(s)ds+Lhe~(s)=ΔNfκ(Un)ΔNfκ(u~(tn))R~hτ(1)(s),s(0,τ),e~(0)=Unu~(tn)=:e~nh0.\begin{split}\left\{\begin{split}&\frac{\,\mathrm{d}\tilde{e}(s)}{\,\mathrm{d}s}+L_{h}\tilde{e}(s)=\Delta_{N}f_{\kappa}(U^{n})-\Delta_{N}f_{\kappa}(\tilde{u}(t_{n}))-\tilde{R}_{h\tau}^{(1)}(s),\quad s\in(0,\tau),\\ &\tilde{e}(0)=U^{n}-\tilde{u}(t_{n})=:\tilde{e}^{n}\in\mathcal{M}_{h}^{0}.\end{split}\right.\end{split} (3.22)

Hence, e~n+1:=e~(τ)\tilde{e}^{n+1}:=\tilde{e}(\tau) satisfies that

e~n+1=ϕ1(τLh)e~n+τϕ0(τLh)[ΔNfκ(Un)ΔNfκ(u~(tn))]0τe~(τs)LhR~hτ(1)(s)ds.\tilde{e}^{n+1}=\phi_{-1}(\tau L_{h})\tilde{e}^{n}+\tau\phi_{0}(\tau L_{h})[\Delta_{N}f_{\kappa}(U^{n})-\Delta_{N}f_{\kappa}(\tilde{u}(t_{n}))]-\int_{0}^{\tau}\tilde{e}^{-(\tau-s)L_{h}}\tilde{R}_{h\tau}^{(1)}(s)\,\mathrm{d}s. (3.23)

Acting I+τLhI+\tau L_{h} on (3.23) and taking the discrete 2\ell^{2} inner product with (ΔN)1e~n+1(-\Delta_{N})^{-1}\tilde{e}^{n+1} yield

e~n+11,N2+τε2Ne~n+1,e~n+1+τκe~n+122=R1,\|\tilde{e}^{n+1}\|_{-1,N}^{2}+\tau\varepsilon^{2}\langle\mathcal{L}_{N}\tilde{e}^{n+1},\tilde{e}^{n+1}\rangle+\tau\kappa\|\tilde{e}^{n+1}\|_{2}^{2}=R_{1}, (3.24)

where

R1=\displaystyle R_{1}= (I+τLh)ϕ1(τLh)e~n,(ΔN)1e~n+1\displaystyle\left\langle(I+\tau L_{h})\phi_{-1}(\tau L_{h})\tilde{e}^{n},(-\Delta_{N})^{-1}\tilde{e}^{n+1}\right\rangle
+τ(I+τLh)ϕ0(τLh)ΔN[fκ(Un)fκ(u~(tn))],(ΔN)1e~n+1\displaystyle+\tau\left\langle(I+\tau L_{h})\phi_{0}(\tau L_{h})\Delta_{N}[f_{\kappa}(U^{n})-f_{\kappa}(\tilde{u}(t_{n}))],(-\Delta_{N})^{-1}\tilde{e}^{n+1}\right\rangle
0τ(I+τLh)e~(τs)LhR~hτ(1)(s),(ΔN)1e~n+1ds.\displaystyle-\int_{0}^{\tau}\left\langle(I+\tau L_{h})\tilde{e}^{-(\tau-s)L_{h}}\tilde{R}_{h\tau}^{(1)}(s),(-\Delta_{N})^{-1}\tilde{e}^{n+1}\right\rangle\,\mathrm{d}s. (3.25)

Next, we use induction argument to prove the convergence. We note that numerical error function e~0=0<12\|\tilde{e}^{0}\|_{\infty}=0<\frac{1}{2} at t=0t=0. Suppose e~n12\|\tilde{e}^{n}\|_{\infty}\leq\frac{1}{2} at tnt_{n}. Thus, the \ell^{\infty} bound for the numerical solutions at tnt_{n} becomes available:

Un=u~(tn)+e~nu~(tn)+e~nuN(tn)+12+12B.\|U^{n}\|_{\infty}=\|\tilde{u}(t_{n})+\tilde{e}^{n}\|_{\infty}\leq\|\tilde{u}(t_{n})\|_{\infty}+\|\tilde{e}^{n}\|_{\infty}\leq\|u_{N}(t_{n})\|_{\infty}+\frac{1}{2}+\frac{1}{2}\leq B.

We now estimate the three terms on the right-hand side given in (3.1.2). For the first linear term, a direct application of Cauchy inequality and Lemma 2.1 give

(I+τLh)ϕ1(τLh)e~n,(ΔN)1e~n+1\displaystyle\left\langle(I+\tau L_{h})\phi_{-1}(\tau L_{h})\tilde{e}^{n},(-\Delta_{N})^{-1}\tilde{e}^{n+1}\right\rangle =(I+τLh)ϕ1(τLh)(ΔN)12e~n,(ΔN)12e~n+1\displaystyle=\left\langle(I+\tau L_{h})\phi_{-1}(\tau L_{h})(-\Delta_{N})^{-\frac{1}{2}}\tilde{e}^{n},(-\Delta_{N})^{-\frac{1}{2}}\tilde{e}^{n+1}\right\rangle
(I+τLh)ϕ1(τLh)(ΔN)12e~n2(ΔN)12e~n+12\displaystyle\leq\|(I+\tau L_{h})\phi_{-1}(\tau L_{h})(-\Delta_{N})^{-\frac{1}{2}}\tilde{e}^{n}\|_{2}\|(-\Delta_{N})^{-\frac{1}{2}}\tilde{e}^{n+1}\|_{2}
12e~n1,N2+12e~n+11,N2.\displaystyle\leq\frac{1}{2}\|\tilde{e}^{n}\|_{-1,N}^{2}+\frac{1}{2}\|\tilde{e}^{n+1}\|_{-1,N}^{2}. (3.26)

By noting that UnB,u~B\|U^{n}\|_{\infty}\leq B,\|\tilde{u}\|_{\infty}\leq B, we have

fκ(Un)fκ(u~(tn))2=fκ(ξn)(Unu~(tn))2fκ(ξn)e~n2|3B2κ1|e~n2:=Ke~n2,\|f_{\kappa}(U^{n})-f_{\kappa}(\tilde{u}(t_{n}))\|_{2}=\|f_{\kappa}^{\prime}(\xi_{n})(U^{n}-\tilde{u}(t_{n}))\|_{2}\leq\|f_{\kappa}^{\prime}(\xi_{n})\|_{\infty}\|\tilde{e}^{n}\|_{2}\leq|3B^{2}-\kappa-1|\|\tilde{e}^{n}\|_{2}:=K\|\tilde{e}^{n}\|_{2},

where ξn\xi_{n} between UnU^{n} and u~(tn)\tilde{u}(t_{n}) and K=|3B2κ1|K=|3B^{2}-\kappa-1|. Then, for the second term in (3.1.2), by using Lemma 2.1 and the above estimate, we have

τ(I+τLh)ϕ0(τLh)ΔN[fκ(Un)fκ(u~(tn))],(ΔN)1e~n+1\displaystyle\tau\left\langle(I+\tau L_{h})\phi_{0}(\tau L_{h})\Delta_{N}[f_{\kappa}(U^{n})-f_{\kappa}(\tilde{u}(t_{n}))],(-\Delta_{N})^{-1}\tilde{e}^{n+1}\right\rangle
τ(I+τLh)ϕ0(τLh)fκ(Un)fκ(u~(tn))2e~n+12\displaystyle\leq\tau\interleave(I+\tau L_{h})\phi_{0}(\tau L_{h})\interleave\|f_{\kappa}(U^{n})-f_{\kappa}(\tilde{u}(t_{n}))\|_{2}\|\tilde{e}^{n+1}\|_{2}
<2τKe~n2e~n+12τKe~n22+τKe~n+122.\displaystyle<2\tau K\|\tilde{e}^{n}\|_{2}\|\tilde{e}^{n+1}\|_{2}\leq\tau K\|\tilde{e}^{n}\|_{2}^{2}+\tau K\|\tilde{e}^{n+1}\|_{2}^{2}. (3.27)

For the last term in (3.1.2), we have the following estimate:

0τ(I+τLh)e~(τs)LhR~hτ(1)(s),(ΔN)1e~n+1ds\displaystyle-\int_{0}^{\tau}\left\langle(I+\tau L_{h})\tilde{e}^{-(\tau-s)L_{h}}\tilde{R}_{h\tau}^{(1)}(s),(-\Delta_{N})^{-1}\tilde{e}^{n+1}\right\rangle\,\mathrm{d}s
0τ(I+τLh)e~(τs)Lhdssupt(0,τ)R~hτ(1)(t)1,Ne~n+11,N\displaystyle\leq\int_{0}^{\tau}\interleave(I+\tau L_{h})\tilde{e}^{-(\tau-s)L_{h}}\interleave\,\mathrm{d}s\sup_{t\in(0,\tau)}\|\tilde{R}_{h\tau}^{(1)}(t)\|_{-1,N}\|\tilde{e}^{n+1}\|_{-1,N}
τ2e~n+11,N2+τ2supt(0,τ)R~hτ(1)(t)1,N2.\displaystyle\leq\frac{\tau}{2}\|\tilde{e}^{n+1}\|_{-1,N}^{2}+\frac{\tau}{2}\sup_{t\in(0,\tau)}\|\tilde{R}_{h\tau}^{(1)}(t)\|_{-1,N}^{2}. (3.28)

For the nonlocal linear term given in (3.24), by appling Lemma 2.2 and Lemma 2.3, we obtain

τε2Ne~n+1,e~n+1\displaystyle-\tau\varepsilon^{2}\langle\mathcal{L}_{N}\tilde{e}^{n+1},\tilde{e}^{n+1}\rangle =τε2(J1)e~n+1Je~n+1,e~n+1\displaystyle=-\tau\varepsilon^{2}\left\langle(J\circledast 1)\tilde{e}^{n+1}-J\circledast\tilde{e}^{n+1},\tilde{e}^{n+1}\right\rangle
=τε2(J1)e~n+122+τε2Je~n+1,e~n+1\displaystyle=-\tau\varepsilon^{2}(J\circledast 1)\|\tilde{e}^{n+1}\|_{2}^{2}+\tau\varepsilon^{2}\langle J\circledast\tilde{e}^{n+1},\tilde{e}^{n+1}\rangle\rangle
=τε2(J1)e~n+122τε2Je~n+1,ΔN((ΔN)1e~n+1)\displaystyle=-\tau\varepsilon^{2}(J\circledast 1)\|\tilde{e}^{n+1}\|_{2}^{2}-\tau\varepsilon^{2}\left\langle J\circledast\tilde{e}^{n+1},\Delta_{N}((-\Delta_{N})^{-1}\tilde{e}^{n+1})\right\rangle
τε2(J1)e~n+122+τγ0e~n+122+τC1γ0e~n+11,N2,\displaystyle\leq-\tau\varepsilon^{2}(J\circledast 1)\|\tilde{e}^{n+1}\|_{2}^{2}+\tau\gamma_{0}\|\tilde{e}^{n+1}\|_{2}^{2}+\tau\frac{C_{1}}{\gamma_{0}}\|\tilde{e}^{n+1}\|_{-1,N}^{2}, (3.29)

where C1C_{1} depends on CJC_{J} and ε\varepsilon.

Therefore, a substitution of (3.1.2)-(3.1.2) and (3.1.2) into (3.24), and recall the define of γ0\gamma_{0}, we get

12(e~n+11,N2e~n1,N2)+(κ+1)τe~n+122\displaystyle\frac{1}{2}(\|\tilde{e}^{n+1}\|_{-1,N}^{2}-\|\tilde{e}^{n}\|_{-1,N}^{2})+(\kappa+1)\tau\|\tilde{e}^{n+1}\|_{2}^{2}
τKe~n22+τKe~n+122+(12+C1γ0)τe~n+11,N2+τ2supt(0,τ)R~hτ(1)(t)1,N2.\displaystyle\leq\tau K\|\tilde{e}^{n}\|_{2}^{2}+\tau K\|\tilde{e}^{n+1}\|_{2}^{2}+(\frac{1}{2}+\frac{C_{1}}{\gamma_{0}})\tau\|\tilde{e}^{n+1}\|_{-1,N}^{2}+\frac{\tau}{2}\sup_{t\in(0,\tau)}\|\tilde{R}_{h\tau}^{(1)}(t)\|_{-1,N}^{2}. (3.30)

Summing the above inequality from 0 to nn leads to

e~n+11,N2+2(κ+1)τk=0n+1e~k22(1+2C1γ0)τk=0n+1e~k1,N2+4τKk=0n+1e~k22+τk=0nsupt(0,τ)R~hτ(1)(t)1,N2.\|\tilde{e}^{n+1}\|_{-1,N}^{2}+2(\kappa+1)\tau\sum_{k=0}^{n+1}\|\tilde{e}^{k}\|_{2}^{2}\leq(1+\frac{2C_{1}}{\gamma_{0}})\tau\sum_{k=0}^{n+1}\|\tilde{e}^{k}\|_{-1,N}^{2}+4\tau K\sum_{k=0}^{n+1}\|\tilde{e}^{k}\|_{2}^{2}+\tau\sum_{k=0}^{n}\sup_{t\in(0,\tau)}\|\tilde{R}_{h\tau}^{(1)}(t)\|_{-1,N}^{2}.

Subsequently, assuming that τ<(1+2C1γ0)1\tau<(1+\frac{2C_{1}}{\gamma_{0}})^{-1} and putting γ1:=2(κ+1)4K0\gamma_{1}:=2(\kappa+1)-4K\geq 0, we can apply the discrete Gronwall’s inequality to obtain the following convergence estimate:

e~n+11,N+(γ1τk=1n+1e~k22)1/2C(τ3+hm),\|\tilde{e}^{n+1}\|_{-1,N}+\left(\gamma_{1}\tau\sum_{k=1}^{n+1}\|\tilde{e}^{k}\|_{2}^{2}\right)^{1/2}\leq C^{*}(\tau^{3}+h^{m}), (3.31)

where CC^{*} depends on C1,γ0,TC_{1},\gamma_{0},T. By using the inverse inequality to (3.31), the following estimate holds

e~n+1\displaystyle\|\tilde{e}^{n+1}\|_{\infty} Ce~n+11,Nh2CC(τ3+hm)h2CC(h3+hm)h2C2Ch3h2=C2Ch12,\displaystyle\leq\frac{C\|\tilde{e}^{n+1}\|_{-1,N}}{h^{2}}\leq\frac{CC^{*}(\tau^{3}+h^{m})}{h^{2}}\leq\frac{C^{\prime}C^{*}(h^{3}+h^{m})}{h^{2}}\leq\frac{C_{2}C^{*}h^{3}}{h^{2}}=C_{2}C^{*}h\leq\frac{1}{2}, (3.32)

provided that h12C2Ch\leq\frac{1}{2C_{2}C^{*}}, where we assumed τCh\tau\leq Ch and the fact that m3m\geq 3. Then, we have

Un+1u~(tn+1)+e~n+1B.\|U^{n+1}\|_{\infty}\leq\|\tilde{u}(t_{n+1})+\tilde{e}^{n+1}\|_{\infty}\leq B. (3.33)

This completes the error estimate for error function e~n+1\tilde{e}^{n+1}.

Now, by the definition of the function u~=uN+τ𝒫Nuτ,1+τ2𝒫Nuτ,2\tilde{u}=u_{N}+\tau\mathcal{P}_{N}u_{\tau,1}+\tau^{2}\mathcal{P}_{N}u_{\tau,2}, we have

en+11,N\displaystyle\|e^{n+1}\|_{-1,N} =Un+1uN(tn+1)1,N\displaystyle=\|U^{n+1}-u_{N}(t_{n+1})\|_{-1,N}
e~n+11,N+u~(tn+1)uN(tn+1)1,N\displaystyle\leq\|\tilde{e}^{n+1}\|_{-1,N}+\|\tilde{u}(t_{n+1})-u_{N}(t_{n+1})\|_{-1,N}
e~n+11,N+τ𝒫N(uτ,1(tn+1))+τ2𝒫N(uτ,2(tn+1))1,N\displaystyle\leq\|\tilde{e}^{n+1}\|_{-1,N}+\|\tau\mathcal{P}_{N}(u_{\tau,1}(t_{n+1}))+\tau^{2}\mathcal{P}_{N}(u_{\tau,2}(t_{n+1}))\|_{-1,N}
e~n+11,N+τ𝒫N(uτ,1(tn+1))1,N+τ2𝒫N(uτ,2(tn+1))1,N.\displaystyle\leq\|\tilde{e}^{n+1}\|_{-1,N}+\tau\|\mathcal{P}_{N}(u_{\tau,1}(t_{n+1}))\|_{-1,N}+\tau^{2}\|\mathcal{P}_{N}(u_{\tau,2}(t_{n+1}))\|_{-1,N}. (3.34)

In a similar manner,

en+12e~n+12+τ𝒫N(uτ,1(tn+1))2+τ2𝒫N(uτ,2(tn+1))2.\|e^{n+1}\|_{2}\leq\|\tilde{e}^{n+1}\|_{2}+\tau\|\mathcal{P}_{N}(u_{\tau,1}(t_{n+1}))\|_{2}+\tau^{2}\|\mathcal{P}_{N}(u_{\tau,2}(t_{n+1}))\|_{2}. (3.35)

Then, thanks to the above observations, we can conclude the error estimate (3.7) from (3.31) and the fact that the boundedness of uτ,1,uτ,2u_{\tau,1},u_{\tau,2}. This competes the proof of Theorem 3.1 (i).

3.2 Convergence analysis of ETD2

Similar to the proof of convergence of ETD1 scheme in subsection 3.1, we next prove the convergence of the ETD2 scheme.

3.2.1 Higher-order consistency analysis

Similar to the higher-order consistency analysis of the ETD1 scheme, the Fourier projection solution uNu_{N} satisfies the following equation:

duNdt+LhuN=(1+ttnτ)ΔNfκ(uN(tn))ttnτΔNfκ(uN(tn1))+τ2g3+𝒪(τ3+hm),t(tn,tn+1],\frac{\,\mathrm{d}u_{N}}{\,\mathrm{d}t}+L_{h}u_{N}=\left(1+\frac{t-t_{n}}{\tau}\right)\Delta_{N}f_{\kappa}(u_{N}(t_{n}))-\frac{t-t_{n}}{\tau}\Delta_{N}f_{\kappa}(u_{N}(t_{n-1}))+\tau^{2}g_{3}+\mathcal{O}(\tau^{3}+h^{m}),\quad t\in(t_{n},t_{n+1}], (3.36)

where the function g3g_{3} depends only on fκf_{\kappa} and uNu_{N} and we have g3¯=0\overline{g_{3}}=0. Define the auxiliary profile

u~=uN+τ2𝒫Nuτ,3,\tilde{u}=u_{N}+\tau^{2}\mathcal{P}_{N}u_{\tau,3}, (3.37)

where the temporal correction function uτ,3u_{\tau,3} solves the following linear ODE system:

duτ,3dt+Lhuτ,3=\displaystyle\frac{\,\mathrm{d}u_{\tau,3}}{\,\mathrm{d}t}+L_{h}u_{\tau,3}= (1+ttnτ)ΔN[fκ(uN(tn))uτ,3(tn)]\displaystyle\left(1+\frac{t-t_{n}}{\tau}\right)\Delta_{N}[f_{\kappa}^{\prime}(u_{N}(t_{n}))u_{\tau,3}(t_{n})]
ttnτΔN[fκ(uN(tn1))uτ,3(tn1)]g3,t(tn,tn+1],\displaystyle-\frac{t-t_{n}}{\tau}\Delta_{N}[f_{\kappa}^{\prime}(u_{N}(t_{n-1}))u_{\tau,3}(t_{n-1})]-g_{3},\quad t\in(t_{n},t_{n+1}], (3.38)

subject to the zero initial value. Note that the solution of (3.2.1) exists and is unique and the solution is smooth enough. Combination of (3.36) and (3.2.1) gives the higher-order consistency estimate:

du~dt+Lhu~=(1+ttnτ)ΔNfκ(u~(tn))ttnτΔNfκ(u~(tn1))+𝒪(τ3+hm),t(tn,tn+1],\frac{\,\mathrm{d}\tilde{u}}{\,\mathrm{d}t}+L_{h}\tilde{u}=\left(1+\frac{t-t_{n}}{\tau}\right)\Delta_{N}f_{\kappa}(\tilde{u}(t_{n}))-\frac{t-t_{n}}{\tau}\Delta_{N}f_{\kappa}(\tilde{u}(t_{n-1}))+\mathcal{O}(\tau^{3}+h^{m}),\quad t\in(t_{n},t_{n+1}], (3.39)

where we have used the estimate

fκ(u~(tk))\displaystyle f_{\kappa}(\tilde{u}(t_{k})) =fκ(uN(tk))+τ2fκ(uN(tk))𝒫Nuτ,3(tk)+𝒪(τ4)\displaystyle=f_{\kappa}(u_{N}(t_{k}))+\tau^{2}f_{\kappa}^{\prime}(u_{N}(t_{k}))\mathcal{P}_{N}u_{\tau,3}(t_{k})+\mathcal{O}(\tau^{4})
=fκ(uN(tk))+τ2𝒫N(fκ(uN(tk))𝒫Nuτ,3(tk))+𝒪(τ4+hm),k=n,n1.\displaystyle=f_{\kappa}(u_{N}(t_{k}))+\tau^{2}\mathcal{P}_{N}(f_{\kappa}^{\prime}(u_{N}(t_{k}))\mathcal{P}_{N}u_{\tau,3}(t_{k}))+\mathcal{O}(\tau^{4}+h^{m}),\quad k=n,n-1.

We also have u~¯=0\overline{\tilde{u}}=0 and the boundedness of u~\|\tilde{u}\|_{\infty}. In fact, by noting uτ,3C\|u_{\tau,3}\|_{\infty}\leq C and if τ\tau is small sufficiently, we have u~uN+u~uNuN+12B.\|\tilde{u}\|_{\infty}\leq\|u_{N}\|_{\infty}+\|\tilde{u}-u_{N}\|_{\infty}\leq\|u_{N}\|_{\infty}+\frac{1}{2}\leq B. And if we adopt the RK2 numerical algorithm for the estimate of U1U^{1}, it holds that U1u~(t1)=𝒪(τ3+hm).U^{1}-\tilde{u}(t_{1})=\mathcal{O}(\tau^{3}+h^{m}).

3.2.2 Convergence analysis in (0,T;Hh1)2(0,T,2)\ell^{\infty}(0,T;H_{h}^{-1})\cap\ell^{2}(0,T,\ell^{2})

For given Un,Un1h0U^{n},U^{n-1}\in\mathcal{M}_{h}^{0}, the solution Un+1U^{n+1} of the ETD2 scheme (2.5) is actually given by Un+1=W(τ)U^{n+1}=W(\tau) with the function W:[0,τ]h0W:[0,\tau]\to\mathcal{M}_{h}^{0} determined by the evolution equation:

{dW(s)ds+LhW(s)=(1+sτ)ΔNfκ(Un)sτΔNfκ(Un1),s(0,τ),W(0)=Un.\begin{split}\left\{\begin{split}&\frac{\,\mathrm{d}W(s)}{\,\mathrm{d}s}+L_{h}W(s)=\left(1+\frac{s}{\tau}\right)\Delta_{N}f_{\kappa}(U^{n})-\frac{s}{\tau}\Delta_{N}f_{\kappa}(U^{n-1}),\quad s\in(0,\tau),\\ &W(0)=U^{n}.\end{split}\right.\end{split} (3.40)

The function W~(s)=u~(tn+s),s[0,τ]\tilde{W}(s)=\tilde{u}(t_{n}+s),\;s\in[0,\tau] satisfies (3.40) with the truncated error R~hτ(2)(s)\tilde{R}_{h\tau}^{(2)}(s)

{dW~(s)ds+LhW~(s)=(1+sτ)ΔNfκ(u~(tn))sτΔNfκ(u~(tn1))+R~hτ(2)(s),s(0,τ),W~(0)=u~(tn),\begin{split}\left\{\begin{split}&\frac{\,\mathrm{d}\tilde{W}(s)}{\,\mathrm{d}s}+L_{h}\tilde{W}(s)=\left(1+\frac{s}{\tau}\right)\Delta_{N}f_{\kappa}(\tilde{u}(t_{n}))-\frac{s}{\tau}\Delta_{N}f_{\kappa}(\tilde{u}(t_{n-1}))+\tilde{R}_{h\tau}^{(2)}(s),\quad s\in(0,\tau),\\ &\tilde{W}(0)=\tilde{u}(t_{n}),\end{split}\right.\end{split} (3.41)

where R~hτ(2)(s)\tilde{R}_{h\tau}^{(2)}(s) satisfies

sups(0,τ)R~hτ(2)(s)1,NC(τ3+hm).\sup\limits_{s\in(0,\tau)}\|\tilde{R}_{h\tau}^{(2)}(s)\|_{-1,N}\leq C(\tau^{3}+h^{m}).

Let e~(s):=W(s)W~(s)\tilde{e}(s):=W(s)-\tilde{W}(s). The difference between (3.40) and (3.41) gives

{de~(s)ds+Lhe~(s)=(1+sτ)[ΔNfκ(Un)ΔNfκ(u~(tn))]sτ[ΔNfκ(Un1)ΔNfκ(u~(tn1))]R~hτ(2)(s),s(0,τ),e~(0)=Unu~(tn):=e~nh0.\begin{split}\left\{\begin{split}&\frac{\,\mathrm{d}\tilde{e}(s)}{\,\mathrm{d}s}+L_{h}\tilde{e}(s)=\left(1+\frac{s}{\tau}\right)[\Delta_{N}f_{\kappa}(U^{n})-\Delta_{N}f_{\kappa}(\tilde{u}(t_{n}))]\\ &\hskip 73.97733pt-\frac{s}{\tau}[\Delta_{N}f_{\kappa}(U^{n-1})-\Delta_{N}f_{\kappa}(\tilde{u}(t_{n-1}))]-\tilde{R}_{h\tau}^{(2)}(s),\quad s\in(0,\tau),\\ &\tilde{e}(0)=U^{n}-\tilde{u}(t_{n}):=\tilde{e}^{n}\in\mathcal{M}_{h}^{0}.\end{split}\right.\end{split} (3.42)

Thus, the solution e~n+1:=e~(τ)\tilde{e}^{n+1}:=\tilde{e}(\tau) satisfies that

e~n+1=\displaystyle\tilde{e}^{n+1}= ϕ1(τLh)e~n+τ(ϕ0+ϕ1)(τLh)[ΔNfκ(Un)ΔNfκ(u~(tn))]\displaystyle\phi_{-1}(\tau L_{h})\tilde{e}^{n}+\tau(\phi_{0}+\phi_{1})(\tau L_{h})\left[\Delta_{N}f_{\kappa}(U^{n})-\Delta_{N}f_{\kappa}(\tilde{u}(t_{n}))\right]
τϕ1(τLh)[ΔNfκ(Un1)ΔNfκ(u~(tn1))]0τe~(τs)LhR~hτ(2)(s)ds.\displaystyle-\tau\phi_{1}(\tau L_{h})\left[\Delta_{N}f_{\kappa}(U^{n-1})-\Delta_{N}f_{\kappa}(\tilde{u}(t_{n-1}))\right]-\int_{0}^{\tau}\tilde{e}^{-(\tau-s)L_{h}}\tilde{R}_{h\tau}^{(2)}(s)\,\mathrm{d}s. (3.43)

Acting I+τLhI+\tau L_{h} on both sides of (3.2.2) and taking 2\ell^{2} inner product of the resulted equation with (ΔN)1e~n+1(-\Delta_{N})^{-1}\tilde{e}^{n+1} yield

e~n+11,N2+τε2Ne~n+1,e~n+1+τκe~n+122=R2,\|\tilde{e}^{n+1}\|_{-1,N}^{2}+\tau\varepsilon^{2}\langle\mathcal{L}_{N}\tilde{e}^{n+1},\tilde{e}^{n+1}\rangle+\tau\kappa\|\tilde{e}^{n+1}\|_{2}^{2}=R_{2}, (3.44)

where

R2=\displaystyle R_{2}= (I+τLh)ϕ1(τLh)e~n,(ΔN)1e~n+1\displaystyle\left\langle(I+\tau L_{h})\phi_{-1}(\tau L_{h})\tilde{e}^{n},(-\Delta_{N})^{-1}\tilde{e}^{n+1}\right\rangle
+τ(I+τLh)(ϕ0+ϕ1)(τLh)ΔN[fκ(Un)fκ(u~(tn))],(ΔN)1e~n+1\displaystyle+\tau\left\langle(I+\tau L_{h})(\phi_{0}+\phi_{1})(\tau L_{h})\Delta_{N}[f_{\kappa}(U^{n})-f_{\kappa}(\tilde{u}(t_{n}))],(-\Delta_{N})^{-1}\tilde{e}^{n+1}\right\rangle
τ(I+τLh)ϕ1(τLh)ΔN[fκ(Un1)fκ(u~(tn1))],(ΔN)1e~n+1\displaystyle-\tau\left\langle(I+\tau L_{h})\phi_{1}(\tau L_{h})\Delta_{N}[f_{\kappa}(U^{n-1})-f_{\kappa}(\tilde{u}(t_{n-1}))],(-\Delta_{N})^{-1}\tilde{e}^{n+1}\right\rangle
0τ(I+τLh)e~(τs)LhR~hτ(2)(s),(ΔN)1e~n+1ds.\displaystyle-\int_{0}^{\tau}\left\langle(I+\tau L_{h})\tilde{e}^{-(\tau-s)L_{h}}\tilde{R}_{h\tau}^{(2)}(s),(-\Delta_{N})^{-1}\tilde{e}^{n+1}\right\rangle\,\mathrm{d}s.

By induction argument, assuming for the numerical error function at the previous time steps tn1t_{n-1} and tnt_{n} satisfy

e~k12,(k=n,n1),\|\tilde{e}^{k}\|_{\infty}\leq\frac{1}{2},\quad(k=n,n-1), (3.45)

then we have Uk=u~(tk)+e~ku~(tk)+e~kuN(tk)+12+12B,(k=n,n1).\|U^{k}\|_{\infty}=\|\tilde{u}(t_{k})+\tilde{e}^{k}\|_{\infty}\leq\|\tilde{u}(t_{k})\|_{\infty}+\|\tilde{e}^{k}\|_{\infty}\leq\|u_{N}(t_{k})\|_{\infty}+\frac{1}{2}+\frac{1}{2}\leq B,(k=n,n-1). Based on the above assumptions and u~B\|\tilde{u}\|_{\infty}\leq B, we have

fκ(Uk)fκ(u~(tk))2|3B2κ1|e~k2:=Ke~k2,k=n,n1.\|f_{\kappa}(U^{k})-f_{\kappa}(\tilde{u}(t_{k}))\|_{2}\leq|3B^{2}-\kappa-1|\|\tilde{e}^{k}\|_{2}:=K\|\tilde{e}^{k}\|_{2},\quad k=n,n-1.

We now estimate R2R_{2} defined in (3.44). By Lemma 2.1 and Cauchy inequality, we obtain

R2\displaystyle R_{2}\leq (I+τLh)ϕ1(τLh)e~n1,Ne~n+11,N\displaystyle\interleave(I+\tau L_{h})\phi_{-1}(\tau L_{h})\interleave\|\tilde{e}^{n}\|_{-1,N}\|\tilde{e}^{n+1}\|_{-1,N}
+τ(I+τLh)(ϕ0+ϕ1)(τLh)fκ(Un)fκ(u~(tn))2e~n+12\displaystyle+\tau\interleave(I+\tau L_{h})(\phi_{0}+\phi_{1})(\tau L_{h})\interleave\|f_{\kappa}(U^{n})-f_{\kappa}(\tilde{u}(t_{n}))\|_{2}\|\tilde{e}^{n+1}\|_{2}
+τ(I+τLh)ϕ1(τLh)fκ(Un1)fκ(u~(tn1))2e~n+12\displaystyle+\tau\interleave(I+\tau L_{h})\phi_{1}(\tau L_{h})\interleave\|f_{\kappa}(U^{n-1})-f_{\kappa}(\tilde{u}(t_{n-1}))\|_{2}\|\tilde{e}^{n+1}\|_{2}
+0τ(I+τLh)e~(τs)Lhdssupt(0,τ)R~hτ(2)(t)1,Ne~n+11,N\displaystyle+\int_{0}^{\tau}\interleave(I+\tau L_{h})\tilde{e}^{-(\tau-s)L_{h}}\interleave\,\mathrm{d}s\sup_{t\in(0,\tau)}\|\tilde{R}_{h\tau}^{(2)}(t)\|_{-1,N}\|\tilde{e}^{n+1}\|_{-1,N}
\displaystyle\leq 12e~n1,N2+12e~n+11,N2+32Kτ(e~n22+e~n+122)\displaystyle\frac{1}{2}\|\tilde{e}^{n}\|_{-1,N}^{2}+\frac{1}{2}\|\tilde{e}^{n+1}\|_{-1,N}^{2}+\frac{3}{2}K\tau(\|\tilde{e}^{n}\|_{2}^{2}+\|\tilde{e}^{n+1}\|_{2}^{2})
+12Kτ(e~n122+e~n+122)+τ2supt(0,τ)R~hτ(2)(t)1,N2+τ2e~n+11,N2.\displaystyle+\frac{1}{2}K\tau(\|\tilde{e}^{n-1}\|_{2}^{2}+\|\tilde{e}^{n+1}\|_{2}^{2})+\frac{\tau}{2}\sup_{t\in(0,\tau)}\|\tilde{R}_{h\tau}^{(2)}(t)\|_{-1,N}^{2}+\frac{\tau}{2}\|\tilde{e}^{n+1}\|_{-1,N}^{2}. (3.46)

Then, a substitution (3.2.2) and (3.1.2) to (3.44), we obtain

12(e~n+11,N2e~n1,N2)+(κ+1)τe~n+122\displaystyle\frac{1}{2}(\|\tilde{e}^{n+1}\|_{-1,N}^{2}-\|\tilde{e}^{n}\|_{-1,N}^{2})+(\kappa+1)\tau\|\tilde{e}^{n+1}\|_{2}^{2} 2Kτe~n+122+32Kτe~n22+12Kτe~n122\displaystyle\leq 2K\tau\|\tilde{e}^{n+1}\|_{2}^{2}+\frac{3}{2}K\tau\|\tilde{e}^{n}\|_{2}^{2}+\frac{1}{2}K\tau\|\tilde{e}^{n-1}\|_{2}^{2}
+(C1γ0+12)τe~n+11,N2+τ2supt(0,τ)R~hτ(2)(t)1,N2.\displaystyle+(\frac{C_{1}}{\gamma_{0}}+\frac{1}{2})\tau\|\tilde{e}^{n+1}\|_{-1,N}^{2}+\frac{\tau}{2}\sup_{t\in(0,\tau)}\|\tilde{R}_{h\tau}^{(2)}(t)\|_{-1,N}^{2}. (3.47)

Summing the above inequality from 1 to nn leads to

e~n+11,N2+2(κ+1)τk=1n+1e~k22\displaystyle\|\tilde{e}^{n+1}\|_{-1,N}^{2}+2(\kappa+1)\tau\sum_{k=1}^{n+1}\|\tilde{e}^{k}\|_{2}^{2}
e~11,N2+(2C1γ0+1)τk=1n+1e~k1,N2+8Kτk=1n+1e~k22+τk=1nsupt(0,τ)R~hτ(2)(t)1,N2.\displaystyle\leq\|\tilde{e}^{1}\|_{-1,N}^{2}+(\frac{2C_{1}}{\gamma_{0}}+1)\tau\sum_{k=1}^{n+1}\|\tilde{e}^{k}\|_{-1,N}^{2}+8K\tau\sum_{k=1}^{n+1}\|\tilde{e}^{k}\|_{2}^{2}+\tau\sum_{k=1}^{n}\sup_{t\in(0,\tau)}\|\tilde{R}_{h\tau}^{(2)}(t)\|_{-1,N}^{2}. (3.48)

Then using the estimate e~11,NC(τ3+hm)\|\tilde{e}^{1}\|_{-1,N}\leq C(\tau^{3}+h^{m}) and assuming that τ<(2C1γ0+1)1\tau<(\frac{2C_{1}}{\gamma_{0}}+1)^{-1} and γ2:=2(κ+1)8K0\gamma_{2}:=2(\kappa+1)-8K\geq 0, an application of the discrete Gronwall’s inequality gives the convergence result:

e~n+11,N+(γ2τk=1n+1e~k22)1/2C2(τ3+hm),\|\tilde{e}^{n+1}\|_{-1,N}+\left(\gamma_{2}\tau\sum_{k=1}^{n+1}\|\tilde{e}^{k}\|_{2}^{2}\right)^{1/2}\leq C_{2}(\tau^{3}+h^{m}), (3.49)

where C2C_{2} is depends on C1,γ0C_{1},\gamma_{0} and TT but independent of τ\tau and hh. Similar to the estimation of (3.32) and (3.33), we can get e~n+112\|\tilde{e}^{n+1}\|_{\infty}\leq\frac{1}{2} and Un+1B\|U^{n+1}\|_{\infty}\leq B provided that τCh,m3\tau\leq Ch,m\geq 3.

Finally, the error estimate (3.8) can be concluded from (3.37), (3.49) and the uniform boundedness of uτ,3u_{\tau,3}. This competes the proof of Theorem 3.1 (ii).

4 Numerical experiments

In this section, we present numerical experiments to verify the temporal convergence rates in the discrete H1H^{-1} norm of the proposed ETD1 and ETD2 schemes. We also investigate the phase transition till the steady state and simulate the coarsening dynamics to show the longtime behavior for the NCH equation (1.1). Set κ=2\kappa=2 for the ETD1 scheme and κ=3\kappa=3 for the ETD2 scheme in all experiments. We use the Guass-type function [9, 31]

J(x)=4πd/2δd+2e|x|2δ2,xd,J(x)=\frac{4}{\pi^{d/2}\delta^{d+2}}e^{-\frac{|x|^{2}}{\delta^{2}}},\quad x\in\mathbb{R}^{d},

where δ>0\delta>0 is a parameter. Note that J1=4/δ2J*1=4/\delta^{2}, then γ0:=ϵ2(J1)1>0\gamma_{0}:=\epsilon^{2}(J\ast 1)-1>0 is equivalent to δ<2ε\delta<2\varepsilon.

4.1 Convergence tests

Example 4.1 (2D test).

We consider the NCH equation (1.1) on Ω=(1,1)2\Omega=(-1,1)^{2} subject to the periodic boundary condition with the initial value u0(x,y)=0.05sinπxsinπyu_{0}(x,y)=0.05\sin\pi x\sin\pi y.

In order to test the time accuracy of the numerical methods, we calculate the H1H^{-1}-norm error using

e(τ):=[h2i,j=0N1((ΔN)12(Ui,jNt(τ,h)Ui,j2Nt(τ/2,h)))]12e(\tau):=\left[h^{2}\sum_{i,j=0}^{N-1}\left((-\Delta_{N})^{-\frac{1}{2}}\left(U_{i,j}^{N_{t}}(\tau,h)-U_{i,j}^{2N_{t}}(\tau/2,h)\right)\right)\right]^{\frac{1}{2}}

and the time convergence order by Rate=ln[e(τ)/e(τ/2)]/ln2\textrm{Rate}=\ln[e(\tau)/e(\tau/2)]/\ln 2.

In this simulation N=256N=256 and T=0.5T=0.5. We compute the numerical solutions for ETD1 and ETD2 schemes with various time step sizes τ=0.005×2k,k=0,1,,8\tau=0.005\times 2^{-k},k=0,1,\cdots,8. The reference solutions is taken as the approximated solution obtained with a quite small time step size τ=0.001×28\tau=0.001\times 2^{-8}. The errors and convergence rates in the discrete H1H^{-1} norm are showed in Table 1 and Table 2, which shows the first order of accuracy for the ETD1 scheme and the second order of accuracy for the ETD2 scheme with different δ\delta and ε\varepsilon, as expected.

Table 1: The H1H^{-1} errors and convergence rates of the ETD1 scheme at time T=0.5T=0.5.
ETD1 δ2=ε2=0.1\delta^{2}=\varepsilon^{2}=0.1 δ2=ε2=0.01\delta^{2}=\varepsilon^{2}=0.01
τ=0.005\tau=0.005 H1H^{-1} error Rate H1H^{-1} error Rate
τ\tau 5.1108E-05 7.2409E-05
τ/2\tau/2 2.1838E-05 1.2267 2.1994E-05 1.7191
τ/4\tau/4 1.0061E-05 1.1181 8.7574E-06 1.3285
τ/8\tau/8 4.8154E-06 1.0630 3.9203E-06 1.1595
τ/16\tau/16 2.3443E-06 1.0385 1.8485E-06 1.0846
τ/32\tau/32 1.1456E-06 1.0331 8.8928E-07 1.0556
τ/64\tau/64 5.5526E-07 1.0448 4.2773E-07 1.0560
τ/128\tau/128 2.6235E-07 1.0817 2.0132E-07 1.0872
τ/256\tau/256 1.1645E-07 1.1718 8.9191E-08 1.1745
Table 2: The H1H^{-1} errors and convergence rates of the ETD2 scheme at time T=0.5T=0.5.
ETD2 δ2=ε2=0.1\delta^{2}=\varepsilon^{2}=0.1 δ2=ε2=0.01\delta^{2}=\varepsilon^{2}=0.01
τ=0.005\tau=0.005 H1H^{-1} error Rate H1H^{-1} error Rate
τ\tau 1.1417E-05 1.8032E-06
τ/2\tau/2 2.5795E-06 2.1461 3.9009E-07 2.2087
τ/4\tau/4 6.2025E-07 2.0562 8.8718E-08 2.1365
τ/8\tau/8 1.5266E-07 2.0225 2.0981E-08 2.0801
τ/16\tau/16 3.7907E-08 2.0098 5.0878E-09 2.0040
τ/32\tau/32 9.4434E-09 2.0051 1.2512E-09 2.0238
τ/64\tau/64 2.3530E-09 2.0048 3.0946E-10 2.0155
τ/128\tau/128 5.8343E-10 2.0119 7.6282E-11 2.0203
τ/256\tau/256 1.4140E-10 2.0448 1.8444E-11 2.0482
Example 4.2 (3D test).

In this test, we calculate the errors and convergence rates by ETD2 scheme for the NCH equation on the computational domain (1,1)3×(0,T](-1,1)^{3}\times(0,T]. We choose the same NN and τ\tau as 2D.

We consider the smooth initial condition u0(x,y,z)=0.05sinπxsinπysinπzu_{0}(x,y,z)=0.05\sin\pi x\sin\pi y\sin\pi z, the computational results are presented in Table 3. From Table 3 one can see that time accuracy is second order, which is onsistent with our theoretical predictions.

Table 3: The H1H^{-1} errors and convergence rates of the ETD2 scheme at time T=0.5T=0.5.
ETD2 ε=0.2,δ=0.1\varepsilon=0.2,\delta=0.1 ε=0.2,δ=0.2\varepsilon=0.2,\delta=0.2
τ=0.005\tau=0.005 H1H^{-1} error Rate H1H^{-1} error Rate
τ\tau 1.8586E-05 2.8680E-05
τ/2\tau/2 3.5861E-06 2.3737 2.4139E-06 3.5706
τ/4\tau/4 8.1176E-07 2.1433 2.6908E-07 3.1652
τ/8\tau/8 1.9593E-07 2.0507 4.4968E-08 2.5811
τ/16\tau/16 4.8344E-08 2.0189 9.7783E-09 2.2012
τ/32\tau/32 1.2017E-08 2.0082 2.3470E-09 2.0587
τ/64\tau/64 2.9918E-09 2.0060 5.7920E-10 2.0187
τ/128\tau/128 7.4157E-10 2.0123 1.4335E-10 2.0146
τ/256\tau/256 1.7970E-10 2.0450 3.4806E-11 2.0421

4.2 Interfaces in the steady states

We now simulate the shapes of the interfaces formed in the steady states by the NCH equation (1.1) under the ETD2 scheme in one-dimensional case with various ε\varepsilon and δ\delta.

Example 4.3 (1D problem).

Let Ω=(1,1)\Omega=(-1,1), u0(x)=0.1(sin2πx+sin3πx)u_{0}(x)=0.1(\sin 2\pi x+\sin 3\pi x), N=1024N=1024 and τ=0.0001\tau=0.0001.

In Figure 2, we fix δ=0.1\delta=0.1 and choose ε\varepsilon as 0.25, 0.2, 0.15, 0.11, 0.1 and 0.09. We plot the numerical solutions at the steady states and the curve of discrete energy evolution. The Figure shows that the numerical solutions reach the steady state at T=10T=10 and T=20T=20, respectively. The energy stability of the ETD2 can be seen in the last column. It is observed that the time required to reach the steady state increases as ε\varepsilon decreases and the interface width becomes sharper for smaller ε\varepsilon.

In Figure 2, we choose ε=0.1\varepsilon=0.1, and δ2\delta^{2} as 0.03, 0.02, 0.01 respectively. Furthermore, we simulate the local CH equation for comparison. From Figure 2, we observe the energy curve remain unchanged after T=10T=10 and the numerical solutions reach the steady state in this time. As δ\delta decreases, the interface turns flat, and the phase transition process is close to the local one for all cases.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
(a) Steady states
Refer to caption
(b) Phases around one interface
Refer to caption
(c) Evolutions of energy
Figure 1: Numerical simulation with δ=0.1\delta=0.1 for different values of ε\varepsilon.
Refer to caption
(a) Steady states
Refer to caption
(b) Phases around one interface
Refer to caption
(c) Evolutions of energy
Figure 2: Numerical simulation with ε=0.1\varepsilon=0.1 for different values of δ\delta.

4.3 Coarsening dynamics and energy evolution

Example 4.4 (2D problem).

We now simulate the long time behavior of the NCH equation (1.1) by using the ETD2 scheme in Ω=(2π,2π)×(2π,2π)\Omega=(-2\pi,2\pi)\times(-2\pi,2\pi) with a random initial data ranging from -0.1 to 0.1. We set time step size τ=0.01\tau=0.01 and choose N=512N=512.

Refer to caption
(a) T=1
Refer to caption
(b) T=10
Refer to caption
(c) T=100
Refer to caption
(d) T=400
Refer to caption
(e) T=1200
Refer to caption
(f) T=2000
Figure 3: Numerical results at T=1,10,100,400,1200,2000T=1,10,100,400,1200,2000.
Refer to caption
(a) δ=0.05,N=512\delta=0.05,N=512
Refer to caption
(b) ε=0.1,N=512\varepsilon=0.1,N=512
Figure 4: The temporal evolution of the energy.

Let ε=δ=0.09\varepsilon=\delta=0.09. Figure 4 shows the coarsening dynamics of numerical solutions, from which one can observe that the dynamic evolves from the initial disorder state to the ordered states and reach the steady state around T=2000T=2000.

Refer to caption
(a) T=0
Refer to caption
(b) T=50
Refer to caption
(c) T=200
Refer to caption
(d) T=500
Refer to caption
(e) T=700
Refer to caption
(f) T=1000
Figure 5: Numerical results at T=0,50,200,500,700,1000T=0,50,200,500,700,1000.
Refer to caption
Figure 6: The temporal evolution of the energy.

Let δ=0.05\delta=0.05 and ε=0.1,0.08,0.06,0.04\varepsilon=0.1,0.08,0.06,0.04, respectively. Figure 4(a) plots the energy evolution curves, and we can observe the energy decay rates comply with the t13t^{-\frac{1}{3}} power law for all cases. This is consistent to the local CH equation [8, 23]. We also plot the curves of the energy for ε=0.1\varepsilon=0.1 with different δ\delta in Figure 4(b). Table 4 presents the linear fitting coefficients mem_{e} and beb_{e} for ε\varepsilon decreasing from 0.10.1 to 0.040.04 with the fitting function E(t)=betmeE(t)=b_{e}t^{m_{e}}.

Table 4: The linear fitting coefficients for the case δ=0.05\delta=0.05.
ε\varepsilon 0.1 0.09 0.08 0.07 0.06 0.05 0.04
mem_{e} -0.314 -0.314 -0.325 -0.331 -0.303 -0.320 -0.341
beb_{e} 21.08 19.49 18.15 17.13 15.04 13.46 12.64
Example 4.5 (3D problem).

We perform the coarsening dynamics of the numerical solutions in 3D on the domain Ω=(2π,2π)3\Omega=(-2\pi,2\pi)^{3} with initial data u0(x,y,z)=0.1+0.2rand(x,y,z)u_{0}(x,y,z)=-0.1+0.2\cdot\text{rand}(x,y,z). Set N=80N=80, τ=0.01\tau=0.01 and ε=δ=0.3\varepsilon=\delta=0.3.

In Figure 6, we plot the contours for numerical solutions at different times. It is shown the numerical solution reaches equilibrium state at around T=1000T=1000 and the energy decay curve is shown in Figure 6.

5 Conclusion remarks

Because of the non-locality and nonlinearity, it is very challenging to prove the convergence of numerical methods for NCH equation. In this paper, we provide a detailed convergence analysis of the ETD1 and ETD2 schemes for the NCH equation, where the Fourier spectral collocation method is used for the spatial discretizations. Due to the lack of the higher-order diffusion term, we adopt (ΔN)1e~n+1(-\Delta_{N})^{-1}\tilde{e}^{n+1} as the test function rather than e~n+1\tilde{e}^{n+1} in the classic CH equation for error equations. The optimal convergence rates in discrete H1H^{-1} norm have been presented by performing the high order consistency analysis. In addition, we also obtain \ell^{\infty} bound of the numerical solutions under some moderate constraints on the space-time step sizes. Finally, we verify the convergence in time of the proposed schemes numerically and simulate the coarsening dynamics to show the long time behavior and present the power law for the energy decay for the NCH equation.

Appendix A: Fourier spectral collocation Approximations

We introduce some notations and useful properties of Fourier spectral collocation approximations for space local linear operators and nonlocal operator in two dimension. Define the index sets

Sh={(i,j)2|1i,jN},S^h={(k,l)2|N2+1i,jN2}.\displaystyle S_{h}=\{(i,j)\in\mathbb{Z}^{2}|1\leq i,j\leq N\},\,\widehat{S}_{h}=\left\{(k,l)\in\mathbb{Z}^{2}|-\frac{N}{2}+1\leq i,j\leq\frac{N}{2}\right\}.

The space of all the Ωh\Omega_{h}-periodic grid functions is denoted by h\mathcal{M}_{h}. For any f,ghf,g\in\mathcal{M}_{h}, the discrete 2\ell^{2} inner product ,\langle\cdot,\cdot\rangle and norm 2\|\cdot\|_{2} and discrete \ell^{\infty} norm \|\cdot\|_{\infty} are, respectively, defined by

f,g=h2(i,j)Shfijgij,f2=f,f,f=max(i,j)Sh|fij|.\langle f,g\rangle=h^{2}\sum_{(i,j)\in S_{h}}f_{ij}g_{ij},\quad\|f\|_{2}=\sqrt{\langle f,f\rangle},\quad\|f\|_{\infty}=\max_{(i,j)\in S_{h}}|f_{ij}|.

For a function fhf\in\mathcal{M}_{h}, we have discrete Fourier transform [29]

fij=(k,l)S^hf^klexp(ikπXxi)exp(ilπXyj),(i,j)Sh.f_{ij}=\sum_{(k,l)\in\hat{S}_{h}}\hat{f}_{kl}\exp(\textrm{i}\frac{k\pi}{X}x_{i})\exp(\textrm{i}\frac{l\pi}{X}y_{j}),\quad(i,j)\in S_{h}. (5.1)

where

f^kl=1N2(i,j)Shfijexp(ikπXxi)exp(ilπXyj),(k,l)S^h,\hat{f}_{kl}=\frac{1}{N^{2}}\sum_{(i,j)\in S_{h}}f_{ij}\exp(-\textrm{i}\frac{k\pi}{X}x_{i})\exp(-\textrm{i}\frac{l\pi}{X}y_{j}),\quad(k,l)\in\widehat{S}_{h}, (5.2)

are the discrete Fourier coefficients. The first and second order derivatives of ff in the xx direction can be defined as

Dxfij=(k,l)S^h(ikπX)f^klexp(ikπXxi)exp(ilπXyj),Dx2fij=(k,l)S^h(ikπX)2f^klexp(ikπXxi)exp(ilπXyj).D_{x}f_{ij}=\sum_{(k,l)\in\hat{S}_{h}}(\textrm{i}\frac{k\pi}{X})\hat{f}_{kl}\exp(\textrm{i}\frac{k\pi}{X}x_{i})\exp(\textrm{i}\frac{l\pi}{X}y_{j}),\quad D_{x}^{2}f_{ij}=\sum_{(k,l)\in\hat{S}_{h}}(\textrm{i}\frac{k\pi}{X})^{2}\hat{f}_{kl}\exp(\textrm{i}\frac{k\pi}{X}x_{i})\exp(\textrm{i}\frac{l\pi}{X}y_{j}).

The differentiation operators in the yy direction can be defined in the same way. For any fhf\in\mathcal{M}_{h}, the discrete gradient and Laplace operators are given by

Nf=(DxfDyf),ΔNf=Dx2f+Dy2f.\nabla_{N}f=\begin{pmatrix}D_{x}f\\ D_{y}f\end{pmatrix},\quad\Delta_{N}f=D_{x}^{2}f+D_{y}^{2}f. (5.3)

For any f,gh,gh×hf,g\in\mathcal{M}_{h},\textbf{{g}}\in\mathcal{M}_{h}\times\mathcal{M}_{h}, we have the following summation-by-parts formulas [22]

f,Ng=Nf,g,f,ΔNg=Nf,Ng=ΔNf,g.\langle f,\nabla_{N}\cdot\textbf{{g}}\rangle=-\langle\nabla_{N}f,\textbf{{g}}\rangle,\quad\langle f,\Delta_{N}g\rangle=-\langle\nabla_{N}f,\nabla_{N}g\rangle=\langle\Delta_{N}f,g\rangle.

For any fhf\in\mathcal{M}_{h}, we call f¯:=14X2f,1\bar{f}:=\frac{1}{4X^{2}}\langle f,1\rangle the mean value of ff. By noticing (1.5) and assuming that the mean of u0u_{0} is zero, we only need to consider the zero-mean grid functions, i.e.,

h0={vh|v,1=0}={vh|v^00=0},\mathcal{M}_{h}^{0}=\{v\in\mathcal{M}_{h}|\langle v,1\rangle=0\}=\{v\in\mathcal{M}_{h}|\hat{v}_{00}=0\},

where v^00=1N2(i,j)Shvij=14X2v,1\hat{v}_{00}=\frac{1}{N^{2}}\sum_{(i,j)\in S_{h}}v_{ij}=\frac{1}{4X^{2}}\langle v,1\rangle. Similar to the continuous case, (ΔN)(-\Delta_{N}) is self-adjoint and positive definite on h0\mathcal{M}_{h}^{0} and thus (ΔN)1(-\Delta_{N})^{-1} is well-defined and is also self-adjoint and positive definite. Then for any f,gh0f,g\in\mathcal{M}_{h}^{0}, we can define the discrete Hh1H_{h}^{-1} inner produce and the discrete Hh1H_{h}^{-1} norm as

f,g1,N:=f,(ΔN)1g=(ΔN)12f,(ΔN)12g,f1,N:=f,f1,N=(ΔN)12f2.\langle f,g\rangle_{-1,N}:=\langle f,(-\Delta_{N})^{-1}g\rangle=\langle(-\Delta_{N})^{-\frac{1}{2}}f,(-\Delta_{N})^{-\frac{1}{2}}g\rangle,\quad\|f\|_{-1,N}:=\sqrt{\langle f,f\rangle_{-1,N}}=\|(-\Delta_{N})^{-\frac{1}{2}}f\|_{2}.

For any f,ϕhf,\phi\in\mathcal{M}_{h}, the discrete convolution fϕhf\circledast\phi\in\mathcal{M}_{h} can be defined componentwise [9] by

(fϕ)ij=h2(m,n)Shfim,jnϕmn,(i,j)Sh.(f\circledast\phi)_{ij}=h^{2}\sum_{(m,n)\in S_{h}}f_{i-m,j-n}\phi_{mn},\quad(i,j)\in S_{h}.

Given a kernel function JJ satisfying the conditions, then for any fhf\in\mathcal{M}_{h}, the discrete form of nonlocal diffusion operator \mathcal{L} can be defined by N=1^N\mathcal{L}_{N}=\mathcal{F}^{-1}\hat{\mathcal{L}}_{N}\mathcal{F} with

^Nf^kl=λklf^kl,(k,l)S^h\hat{\mathcal{L}}_{N}\hat{f}_{kl}=\lambda_{kl}\hat{f}_{kl},\quad(k,l)\in\hat{S}_{h} (5.4)

where \mathcal{F} is a discrete Fourier transform matrix and

λkl=h2(i,j)ShJ(xi,yj)(1exp(ikπXxi)exp(ilπXyj)).\lambda_{kl}=h^{2}\sum_{(i,j)\in S_{h}}J(x_{i},y_{j})\left(1-\exp(-\textrm{i}\frac{k\pi}{X}x_{i})\exp(-\textrm{i}\frac{l\pi}{X}y_{j})\right).

References

  • AE [04] A. Archer and R. Evans. Dynamical density functional theory and its application to spinodal decomposition. J. Chem. Phys., 121:4246–4254, 2004.
  • Bat [06] P. W. Bates. On some nonlocal evolution equations arising in materials science. Fields Inst. Commun., 48:13–52, 2006.
  • [3] P. W. Bates and J. L. Han. The Dirichlet boundary problem for a nonlocal Cahn-Hilliard equation. J. Math. Anal., 311:289–312, 2005.
  • [4] P. W. Bates and J. L. Han. The Neumann boundary problem for a nonlocal Cahn-Hilliard equation. J. Differ. Equa., 212:235–277, 2005.
  • BKV [98] G. Beylkin, J. M. Keiser, and Lev. Vozovoi. A new class of time discretization schemes for the solution of nonlinear PDEs. J. Comput. Phys., 147(2):362–387, 1998.
  • CH [58] J. W. Cahn and J. E. Hilliard. Free energy of a nonuniform system. I. Interfacial free energy. J. Chem. Phys., 28(2):258–267, 1958.
  • CM [02] S. M. Cox and P. C. Matthews. Exponential time differencing for stiff systems. J. Comput. Phys., 176(2):430–455, 2002.
  • DD [16] S. B. Dai and Q. Du. Computational studies of coarsening rates for the Cahn–Hilliard equation with phase-dependent diffusion mobility. J. Comput. Phys., 310:85–108, 2016.
  • DJLQ [18] Q. Du, L. L. Ju, X. Li, and Z. H. Qiao. Stabilized linear semi-implicit schemes for the nonlocal Cahn–Hilliard equation. J. Comput. Phys., 363:39–54, 2018.
  • DJLQ [19] Q. Du, L. L. Ju, X. Li, and Z. H. Qiao. Maximum principle preserving exponential time differencing schemes for the nonlocal Allen–Cahn equation. SIAM J. Numer. Anal., 57(2):875–898, 2019.
  • DJLQ [21] Q. Du, L. L. Ju, X. Li, and Z. H. Qiao. Maximum bound principles for a class of semilinear parabolic equations and exponential time-differencing schemes. SIAM Review., 63(2):317–359, 2021.
  • DZ [04] Q. Du and W. X. Zhu. Stability analysis and application of the exponential time differencing schemes. J. Comput. Math., 22(2):200–209, 2004.
  • DZ [05] Q. Du and W. X. Zhu. Analysis and applications of the exponential time differencing schemes and their contour integration modifications. BIT Numerical. Mathematics., 45:307–328, 2005.
  • Fif [03] P. Fife. Some Nonclassical Trends in Parabolic and Parabolic-like Evolutions, Trends in Nonlinear Analysis. Springer, Berlin., 2003.
  • FSY [24] Z. H. Fu, J. Shen, and J. Yang. Higher-Order Energy-Decreasing Exponential Time Differencing Runge-Kutta methods for Gradient Flows. arXiv preprint arXiv:2402.15142, 2024.
  • GG [05] H. Gajewski and K. Gärtner. On a nonlocal model of image segmentation. Z. Angew. Math. Phys., 56:572–591, 2005.
  • GLW [17] Z. Guan, J. Lowengrub, and C. Wang. Convergence analysis for second-order accurate schemes for the periodic nonlocal Allen-Cahn and Cahn-Hilliard equations. Math. Methods Appl. Sci., 40(18):6836–6863, 2017.
  • GLWW [14] Z. Guan, J. Lowengrub, C. Wang, and S. M. Wise. Second order convex splitting schemes for periodic nonlocal Cahn–Hilliard and Allen–Cahn equations. J. Comput. Phys., 277:48–71, 2014.
  • GWW [14] Z. Guan, C. Wang, and S. M. Wise. A convergent convex splitting scheme for the periodic nonlocal Cahn-Hilliard equation. Numer Math., 128:377–406, 2014.
  • HKV [01] D. J. Horntrop, M. A. Katsoulakis, and D. G. Vlachos. Spectral methods for mesoscopic models of pattern formation. J. Comput. Phys., 173:364–390, 2001.
  • HO [10] M. Hochbruck and A. Ostermann. Exponential integrators. Acta Numerica., 19:209–286, 2010.
  • JLQZ [18] L. L. Ju, X. Li, Z. H. Qiao, and H. Zhang. Energy stability and error estimates of exponential time differencing schemes for the epitaxial growth model without slope selection. Math Comput., 87(312):1859–1885, 2018.
  • JZD [15] L. L. Ju, J. Zhang, and Q. Du. Fast and accurate algorithms for simulating coarsening dynamics of Cahn–Hilliard equations. Comput. Mater. Sci., 108:272–282, 2015.
  • LJM [19] X. Li, L. L. Ju, and X. C. Meng. Convergence analysis of exponential time differencing schemes for the Cahn-Hilliard equation. Commun. Comput. Phys., 26(5):1510–1529, 2019.
  • LQT [16] D. Li, Z. H. Qiao, and T. Tang. Characterizing the stabilization size for semi-implicit Fourier-spectral method to phase field equations. SIAM J. Numer. Anal., 54(3):1653–1681, 2016.
  • LQW [21] X. Li, Z. H. Qiao, and C. Wang. Convergence analysis for a stabilized linear semi-implicit numerical scheme for the nonlocal Cahn–Hilliard equation. Math Comput., 90(327):171–188, 2021.
  • LQW [23] X. Li, Z. H. Qiao, and C. Wang. Stabilization parameter analysis of a second-order linear numerical scheme for the nonlocal Cahn–Hilliard equation. SIAM J. Numer. Anal., 43(2):1089–1114, 2023.
  • LQW [24] X. Li, Z. H. Qiao, and C. Wang. Double stabilizations and convergence analysis of a second-order linear numerical scheme for the nonlocal Cahn-Hilliard equation. Sci. China. Math., 67(1):187–210, 2024.
  • STW [11] J. Shen, T. Tang, and L-L. Wang. Spectral methods: algorithms, analysis and applications. Springer., 2011.
  • ZS [23] Q. Zhou and Y. B. Sun. Energy stability of exponential time differencing schemes for the nonlocal Cahn-Hilliard equation. Numer. Methods. Partial. Differ. Equ., 39(5):4030–4058, 2023.
  • ZW [24] D. N. Zhang and D. L. Wang. Energy-decreasing second order exponential time differencing Runge–Kutta methods for Nonlocal Cahn–Hilliard equation. Appl. Math. Lett., 150:108974, 2024.