\newtotcounter

citnum \regtotcountertable \regtotcounterfigure

EFFICIENT AND STABLE TIME INTEGRATION OF CAHN-HILLIARD EQUATIONS: EXPLICIT, IMPLICIT AND EXPLICIT ITERATIVE SCHEMES

© 2024 M.A. Botchev,1,*, I.A. Fahurdinov, 1,2,**, E.B. Savenkov1,***

1 Keldysh Institute of Applied Mathematics of Russian Academy of Sciences,
Miusskaya Sq. 4, Moscow, 125047, Russia
2 National Research Nuclear University ‘‘Moscow Engineering Physics Institute’’ (MEPhI)
Kashirskoe Hw. 31, Moscow, 115409, Russia
*e–mail: [email protected]
** [email protected]
*** [email protected]

Submitted –.–.2024

Revised version –.–.2024

Accepted for publication –.–.2024

Abstract

To solve the Cahn-Hilliard equation numerically, a new time integration algorithm is proposed, which is based on a combination of the Eyre splitting and the local iteration modified (LIM) scheme. The latter is employed to tackle the implicit system arising each time integration step. The proposed method is gradient-stable and allows to use large time steps, whereas, regarding its computational structure, it is an explicit time integration scheme. Numerical tests are presented to demonstrate abilities of the new method and to compare it with other time integration methods for Cahn-Hilliard equation. Citations: \totalcitnum. Figures: \totalfigure.

Key words: Cahn-Hilliard equation, gradient-stable schemes, Eyre splitting, local iteration modified scheme, LIM.

DOI:

1.  Introduction

The Cahn-Hilliard equation was proposed in 1958 (see original paper [1]) to describe a phase separation in two-component alloys. Currently the equation is widely applied in various research fields and is one of the fundamental models in the so called gradient or weakly non-local thermomechanics of continuous media [2]. As a part of more complex models, the equation is employed in multiphase hydrodynamics, material science, solidification problems, phase transition theory and in many other fields [3, 4]. From a theoretical point of view, the Cahn-Hilliard equation forms the basis of phenomenological spinodal decomposition theory [5] and has a broad range of applications in theoretical physics, see [6, 7, 8] and references therein. Mathematics behind the Cahn-Hilliard equation is also widely discussed in the literature, see, e.g., [9].

Up to recently, the Cahn-Hilliard equation had been primarily seen as a theoretical tool in mathematical modeling of various processes. However, last decennia it has been actively employed also as a simulation tool, either independently or a part of more complex models. This has necessitated development of efficient numerical algorithms to solve the equation.

Difficulties arising in design of efficient numerical algorithms for the Cahn-Hilliard equation are mainly caused by the following two reasons. First, the Cahn-Hilliard equation is a nonlinear PDE containing fourth order space derivatives. Due to nonlinearities, solutions of the Cahn-Hilliard equation evolve on a broad range of space and time scales. In particular, a typical solution is almost constant within the certain space regions of homogeneity which correspond to ‘‘pure’’ phases of the system. These homogeneity regions have a typical size d𝑑ditalic_d and are separated from each other by thin layers, of width ϵdmuch-less-thanϵ𝑑\epsilonup\ll droman_ϵ ≪ italic_d, the so-called ‘‘diffuse interfaces’’, where the solution varies from its minimal to maximal values but remains smooth. At initial stages of spinodal decomposition, usually emerging due to a perturbation in the initial random distribution, the diffuse interface width ϵϵ\epsilonuproman_ϵ is comparable to homogeneity region size d𝑑ditalic_d (ϵdsimilar-toϵ𝑑\epsilonup\sim droman_ϵ ∼ italic_d) and typical evolution times are of order ϵ2superscriptϵ2\epsilonup^{2}roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. However, in the course of further evolution, when the diffuse interface width becomes considerably smaller than the homogeneity region size, ϵdmuch-less-thanϵ𝑑\epsilonup\ll droman_ϵ ≪ italic_d, typical evolution times are of order 1/ϵ1ϵ1/\epsilonup1 / roman_ϵ, see [10, 11].

As a consequence of these effects, explicit time integration schemes, being applied to the Cahn-Hilliard equation, are stable for time step sizes τh4similar-toτsuperscript4\tauup\sim h^{4}roman_τ ∼ italic_h start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT, with hhitalic_h being the space grid size. Although such a restriction is physically motivated, in some cases it may turn out unacceptably strict for a computationally efficient solution process. At the same time, in fully implicit schemes, uniqueness of the solution on the next time level is not guaranteed for sufficiently large time step sizes [12, 13]. Therefore, design of numerical schemes which would be (a) uniquely solvable for arbitrary time steps, (b) stable, and (c) conservative, is not an easy task. Baseline algorithms meeting these requirements appeared relatively recently [14, 15]. They are based on a special splitting of system energy called convex splitting, see Section 3. We note that, in the context of numerical solution of the Cahn-Hilliard equation, ‘‘stability’’ usually means the so-called ‘‘gradient stability’’ (also known as ‘‘energy stability’’), which guarantees that a discretized analogue of the system free energy does not grow with time, see Sections 23 below. This stability condition can be considered as strong in the sense that the expression for the free energy includes first derivatives of solution. This is necessary to make sure that the numerical solution of the Cahn-Hilliard equation is thermodynamically consistent.

Currently, important research directions in efficient time integration of the Cahn-Hilliard equation include construction of adaptive numerical algorithms (needed due to the broad range of characteristic evolution times, as discussed above), development of energy-stable time integration schemes (producing thermodynamically consistent numerical solutions), and design of schemes which would be, on one hand, sufficiently asymptotically stable and, on the other hand, uniquely solvable and computationally efficient.

A complete review of current research on design and analysis of time integration schemes for the Cahn-Hilliard equation is far beyond the scope of this paper. An overview of main research directions in this field is given in a survey [16]. As examples of other relevant literature, we note [17, 18, 19, 20, 21, 22, 23] (adaptive and high-order methods), [24, 25, 26] (schemes, allowing ‘‘large’’ time step sizes), and [14, 15, 27, 28, 29, 30, 31, 32, 33] (energy-stable schemes).

To discretize the Cahn-Hilliard equation in space, various methods can be employed, such as classic finite difference methods, finite volume methods, spectral methods, standard and isogeometric finite element methods, see [34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47].

The aim of this paper is to study numerically a new class of time integration schemes for the Cahn-Hilliard equation. The new algorithms are based on two main ideas: (a) employment of implicit-explicit time integration schemes based on the convex splitting of the system energy [14, 15], ensuring the scheme is energy-stable, and (b) application of the local iteration modified (LIM) time integration scheme, which allows to combine the computational simplicity of explicit schemes with stability properties of implicit schemes. The Eyre splitting [14, 15] can currently be seen as a basis for constructing time integration schemes for the Cahn-Hilliard equations, which should be both energy stable and uniquely solvable (meaning that the solution on the next time level in the time integration scheme exists and is unique).

Local iteration schemes form a family of explicit time integration methods, where stability is achieved by employing Chebyshev polynomials. Chebyshev polynomial iterations are used in time integration since at least the 1950s [48, 49]. It should be emphasized that Chebyshev iteration in the local iteration schemes are used not to approximately solve linear systems arising in implicit time integration; if this would be the case then, as shown in 1952 by Gel’fand and Lokutsievskii (see [50] and [51, Ch.10, § 4.12]), no essential gain in computational work with respect to explicit schemes can be obtained. The key characteristic feature of the local iteration schemes is that Chebyshev iterations are constructed to ensure stability and accuracy of time integration, rather than to achieve a fast convergence to a solution of an implicit scheme. The local iteration schemes are proposed in [52, 53, 54, 55] and further developed in [56, 57, 58], see also references in [58]. Specially tuned Chebyshev iterations allow the local iteration schemes to work with time step sizes essentially larger than in explicit Euler scheme, while keeping the computational simplicity of explicit schemes.

The first local iteration scheme proposed in [53, 52, 54] is not monotone (its numerical solution is not guaranteed to be nonnegative) and is not asymptotically stable for large times t𝑡t\rightarrow\inftyitalic_t → ∞. Therefore here the local iteration modified (LIM) scheme, proposed in [55], is employed, which does have there properties. A detailed description and comparison of the local iteration schemes can be found in [58]. Up to date, the LIM scheme has been successfully applied to solve various rather challenging real-life problems, see, e.g., [59, 60, 61]. In [62], the LIM scheme is applied to solve the Cahn-Hilliard equation within a numerical implementation of a full mathematical model to describe metal crystallization processes. There, at each time step, the Cahn-Hilliard equation is linearized and the linearized problem is handled by the LIM method. No comparisons with other relevant time integration schemes are carried out in [62], and properties such as gradient stability are not investigated either.

In this paper, with a space one-dimensional Cahn-Hilliard equation taken as example, we propose new time integration schemes which are based the LIM scheme and implicit-explicit Eyre splittings. Results of numerical experiments, where the proposed methods are tested and compared to other time integration schemes, are presented.

In Section 2 a short description of the model is given. Currently used time integration schemes are briefly discussed in Section 3 and then, in the same section, new schemes are presented. Numerical experiments are described in Section 4. Finally, Section 5 summarizes results obtained in this paper.

Work of E.B. Savenkov is supported by by the Russian Science Foundation grant no. 2-11-00203.

2.  Mathematical model

We consider an one-dimensional (1D) Cahn-Hilliard equation

ct=x(Mμx),μ(c)=F(c)ϵ22c2x,F(c)=c2(1c)2,formulae-sequence𝑐𝑡𝑥𝑀μ𝑥formulae-sequenceμ𝑐superscript𝐹𝑐superscriptϵ2superscript2𝑐superscript2𝑥𝐹𝑐superscript𝑐2superscript1𝑐2\frac{\partial c}{\partial t}=\frac{\partial}{\partial x}\left(M\frac{\partial% \muup}{\partial x}\right),\quad\muup(c)=F^{\prime}(c)-\epsilonup^{2}\frac{% \partial^{2}c}{\partial^{2}x},\quad F(c)=c^{2}(1-c)^{2},divide start_ARG ∂ italic_c end_ARG start_ARG ∂ italic_t end_ARG = divide start_ARG ∂ end_ARG start_ARG ∂ italic_x end_ARG ( italic_M divide start_ARG ∂ roman_μ end_ARG start_ARG ∂ italic_x end_ARG ) , roman_μ ( italic_c ) = italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_c ) - roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_c end_ARG start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x end_ARG , italic_F ( italic_c ) = italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - italic_c ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (2.1)

where c=c(x,t)𝑐𝑐𝑥𝑡c=c(x,t)italic_c = italic_c ( italic_x , italic_t ) is the unknown function, M=M(c)>0𝑀𝑀𝑐0M=M(c)>0italic_M = italic_M ( italic_c ) > 0 is a diffusional mobility coefficient (in all tests presented here we set M1𝑀1M\equiv 1italic_M ≡ 1), μ=μ(c,c/x)μμ𝑐𝑐𝑥\muup=\muup(c,\partial{c}/\partial{x})roman_μ = roman_μ ( italic_c , ∂ italic_c / ∂ italic_x ) is a chemical potential, ϵ>0ϵ0\epsilonup>0roman_ϵ > 0, ϵ=constϵconst\epsilonup=\text{const}roman_ϵ = const is the diffuse interface width, t𝑡titalic_t is the time, x𝑥xitalic_x the space variable.

Equation (2.1) is solved in a 1D domain Ω=(0,1)Ω01\Omega=(0,1)roman_Ω = ( 0 , 1 ) for t(0,T]𝑡0𝑇t\in(0,T]italic_t ∈ ( 0 , italic_T ]. Boundary conditions

cx=μx=0,𝑐𝑥μ𝑥0\frac{\partial c}{\partial x}=\frac{\partial\muup}{\partial x}=0,divide start_ARG ∂ italic_c end_ARG start_ARG ∂ italic_x end_ARG = divide start_ARG ∂ roman_μ end_ARG start_ARG ∂ italic_x end_ARG = 0 , (2.2)

are set on the domain boundary and, for  t=0𝑡0t=0italic_t = 0, initial condition is given,

c(x,0)=c0(x).𝑐𝑥0superscript𝑐0𝑥c(x,0)=c^{0}(x).italic_c ( italic_x , 0 ) = italic_c start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ( italic_x ) . (2.3)

From a physical point of view, equation (2.1) can be derived as follows [63]. Consider a two-phase system with component concentrations c1,2subscript𝑐12c_{1,2}italic_c start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT, c1+c2=1subscript𝑐1subscript𝑐21c_{1}+c_{2}=1italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1, and assume that the free energy has a form, with c=c1𝑐subscript𝑐1c=c_{1}italic_c = italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT,

Ψ[c,c]=Ωψ(c,c)𝑑x,Ψ𝑐𝑐subscriptΩψ𝑐𝑐differential-d𝑥\Psi[c,\nabla c]=\int\limits_{\Omega}\psiup(c,\nabla c)\,dx,roman_Ψ [ italic_c , ∇ italic_c ] = ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT roman_ψ ( italic_c , ∇ italic_c ) italic_d italic_x , (2.4)
ψ(c,c)=F(c)+ϵ22|c|2,F(c)=c2(1c)2,formulae-sequenceψ𝑐𝑐𝐹𝑐superscriptϵ22superscript𝑐2𝐹𝑐superscript𝑐2superscript1𝑐2\psiup(c,\nabla c)=F(c)+\frac{\epsilonup^{2}}{2}|\nabla c|^{2},\quad F(c)=c^{2% }(1-c)^{2},roman_ψ ( italic_c , ∇ italic_c ) = italic_F ( italic_c ) + divide start_ARG roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG | ∇ italic_c | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , italic_F ( italic_c ) = italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - italic_c ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (2.5)

where ψψ\psiuproman_ψ is the system free energy density, F𝐹Fitalic_F is the so-called double-well potential, ϵ>0ϵ0\epsilonup>0roman_ϵ > 0 is a small parameter. The first term in (2.5) describes the ‘‘dividing’’ part of the free energy which ensures that the phases do not mix and, hence, the regions with constant c1,2subscript𝑐12c_{1,2}italic_c start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT emerge. The second term in (2.5) allows to relate a given energy to the diffuse interface. Parameter ϵϵ\epsilonuproman_ϵ there defines the width of the diffuse interface separating the ‘‘pure’’ phases. Inclusion of the gradient terms makes the system energy depend not only on the quantity of each of the two components but also on the shape of regions taken up by them. Energy form (2.4), (2.5) arises in various models and is typical for weakly non-local (or gradient) thermomechanics.

It should be emphasized that potential F(c)𝐹𝑐F(c)italic_F ( italic_c ) in (2.5) is empirical. Its characteristic feature is the two minima at c=0𝑐0c=0italic_c = 0 and c=1𝑐1c=1italic_c = 1, corresponding to the ‘‘pure’’ phases. The state c=1/2𝑐12c=1/2italic_c = 1 / 2 corresponding to the maximum of F(c)𝐹𝑐F(c)italic_F ( italic_c ) is unstable and, hence, the mixing of the two phases is prevented by the instability. A more accurate and better physically motivated energy model is the so-called logarithmic potential [63]. It possesses essentially the same characteristic properties as the energy form considered here, and the latter can be seen as an approximation of the former.

An illustration of these observations is given in Figure 1. In the left plot, the ‘‘separating’’ part ψdwFsubscriptψdw𝐹\psiup_{\text{dw}}\equiv Froman_ψ start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT ≡ italic_F of the free energy is shown. The two minima correspond to the ‘‘pure’’ phases, i.e., to the states c=c1=0𝑐subscript𝑐10c=c_{1}=0italic_c = italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0, c2=1subscript𝑐21c_{2}=1italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1 and c=c1=1𝑐subscript𝑐11c=c_{1}=1italic_c = italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 1, c2=0subscript𝑐20c_{2}=0italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0. The state in the neighborhood of c=1/2𝑐12c=1/2italic_c = 1 / 2 is unstable. In the right plot of the figure, a typical stationary solution of the system and energy distribution are shown. As can be seen, the gradient part of the free energy is nonzero only in the diffuse interface regions. At the same time, the separating part of the free energy is zero outside the diffuse interface regions.

Refer to caption
Refer to caption
Figure 1: Left: an example of the ‘‘separating’’ free energy part ψdwsubscriptψdw\psiup_{\text{dw}}roman_ψ start_POSTSUBSCRIPT dw end_POSTSUBSCRIPT. Right: a typical solution and energy distribution.

Furthermore, one can show that if c𝑐citalic_c is a conservative quantity (as is in our case), the kinetic equation describing the field evolution c=c(x,t)𝑐𝑐𝑥𝑡c=c(x,t)italic_c = italic_c ( italic_x , italic_t ) reads [6, 63]:

ct=(Mμ),μ=δψ[c,c]δc,formulae-sequence𝑐𝑡𝑀μμδψ𝑐𝑐δ𝑐\frac{\partial c}{\partial t}=-\nabla\mathbf{\cdot}\left(-M\nabla\muup\right),% \quad\muup=\frac{\deltaup\psiup[c,\nabla c]}{\deltaup c},divide start_ARG ∂ italic_c end_ARG start_ARG ∂ italic_t end_ARG = - ∇ ⋅ ( - italic_M ∇ roman_μ ) , roman_μ = divide start_ARG roman_δ roman_ψ [ italic_c , ∇ italic_c ] end_ARG start_ARG roman_δ italic_c end_ARG , (2.6)

where δ()/δcδδ𝑐\deltaup(\cdot)/\deltaup croman_δ ( ⋅ ) / roman_δ italic_c is the functional derivative (Gateaux derivative). For M𝑀Mitalic_M constant, due to (2.5), the last relation takes form

1Mct=Δμ,μ=ϵ2Δc+F(c),formulae-sequence1𝑀𝑐𝑡Δμμsuperscriptϵ2Δ𝑐superscript𝐹𝑐\frac{1}{M}\frac{\partial c}{\partial t}=\Delta\muup,\quad\muup=-\epsilonup^{2% }\Delta{c}+F^{\prime}(c),divide start_ARG 1 end_ARG start_ARG italic_M end_ARG divide start_ARG ∂ italic_c end_ARG start_ARG ∂ italic_t end_ARG = roman_Δ roman_μ , roman_μ = - roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ italic_c + italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_c ) ,

where ΔΔ\Deltaroman_Δ is the Laplacian. In the 1D case, under consideration here, this coincides with (2.1).

As can also be shown [12, 63], equation (2.1) has solutions c=c(x,t)𝑐𝑐𝑥𝑡c=c(x,t)italic_c = italic_c ( italic_x , italic_t ) with the following fundamental properties:

ddtΩc=0,ddtΨ[c,c]0,formulae-sequence𝑑𝑑𝑡subscriptΩ𝑐0𝑑𝑑𝑡Ψ𝑐𝑐0\frac{d}{dt}\int\limits_{\Omega}c=0,\qquad\frac{d}{dt}\Psi[c,\nabla c]% \leqslant 0,divide start_ARG italic_d end_ARG start_ARG italic_d italic_t end_ARG ∫ start_POSTSUBSCRIPT roman_Ω end_POSTSUBSCRIPT italic_c = 0 , divide start_ARG italic_d end_ARG start_ARG italic_d italic_t end_ARG roman_Ψ [ italic_c , ∇ italic_c ] ⩽ 0 ,

which means that c𝑐citalic_c is conserved and the free energy is a non-increasing function of c𝑐citalic_c being the solution of (2.1) or (2.6) and satisfying homogeneous boundary conditions (2.2).

These properties of the analytical solution immediately put restrictions on a numerical scheme for solving (2.1): it should be conservative (which is relatively easy to achieve) and energy-stable, i.e., delivering a discrete analogue of the last inequality (which is more difficult to achieve). In addition, the discrete problem should be uniquely solvable for a given time step size. In practice, this may turn out to be a rather nontrivial task because the potential F(c)𝐹𝑐F(c)italic_F ( italic_c ) is not a convex function and, hence, a discrete solution may be found in a local minimum which is not global and not achievable from an initial state.

3.  Numerical algorithms

This section deals with numerical algorithms applied to solve the Cahn-Hilliard equation. In the first subsection classical time integration schemes are considered, having various degree of implicitness. The second subsection describes the Eyre splitting for the time integration of the Cahn-Hilliard equations. New algorithms based on the LIM scheme are presented in the third subsection.

3.1.  Classical time integration schemes

We now consider a number of time integration schemes taking part in the numerical comparisons presented below. All these schemes are well known and described in numerous literature and, in particular, in [64]. All schemes considered in this subsection approximate equation (2.1) and are conservative.

Throughout this paper space derivatives are discretized by finite differences. Assume that a uniform grid with grid size h=1/N1𝑁h=1/Nitalic_h = 1 / italic_N and nodes xi=(i1/2)hsubscript𝑥𝑖𝑖12x_{i}=(i-1/2)hitalic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ( italic_i - 1 / 2 ) italic_h, i=0,N+1¯𝑖¯0𝑁1i=\overline{0,N+1}italic_i = over¯ start_ARG 0 , italic_N + 1 end_ARG, with N𝑁Nitalic_N being the number of grid cells, is used to discretize the equation in the domain ΩΩ\Omegaroman_Ω. The domain boundaries x=0𝑥0x=0italic_x = 0 and x=1𝑥1x=1italic_x = 1 correspond to points x1/2=(x0+x1)/2=0subscript𝑥12subscript𝑥0subscript𝑥120x_{1/2}=(x_{0}+x_{1})/2=0italic_x start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT = ( italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) / 2 = 0 and xN+1/2=(xN+xN+1)/2=Nhsubscript𝑥𝑁12subscript𝑥𝑁subscript𝑥𝑁12𝑁x_{N+1/2}=(x_{N}+x_{N+1})/2=Nhitalic_x start_POSTSUBSCRIPT italic_N + 1 / 2 end_POSTSUBSCRIPT = ( italic_x start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT + italic_x start_POSTSUBSCRIPT italic_N + 1 end_POSTSUBSCRIPT ) / 2 = italic_N italic_h and the grid nodes x1,x2,,xNsubscript𝑥1subscript𝑥2subscript𝑥𝑁x_{1},x_{2},\ldots,x_{N}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_x start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT lie within the domain. For time discretization a uniform grid with time step size ττ\tauuproman_τ and nodes tn=nτsubscript𝑡𝑛𝑛τt_{n}=n\tauupitalic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = italic_n roman_τ, n=0,1,2,𝑛012n=0,1,2,\ldotsitalic_n = 0 , 1 , 2 , …, is used. Thus, the solution of the discretized problem is defined in points (xi,tj)subscript𝑥𝑖subscript𝑡𝑗(x_{i},t_{j})( italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ), i=1,N¯𝑖¯1𝑁i=\overline{1,N}italic_i = over¯ start_ARG 1 , italic_N end_ARG, j=0,1,𝑗01j=0,1,\ldotsitalic_j = 0 , 1 , …. The corresponding values of the grid functions chsubscript𝑐c_{h}italic_c start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT are denoted by cin=c(xi,tn)superscriptsubscript𝑐𝑖𝑛𝑐subscript𝑥𝑖subscript𝑡𝑛c_{i}^{n}=c(x_{i},t_{n})italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = italic_c ( italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ).

Let ΔhsubscriptΔ\Delta_{h}roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT be a standard three-point finite-difference approximation of the Laplacian. In the 1D case under consideration, its values in the nodes i=1,N1¯𝑖¯1𝑁1i=\overline{1,N-1}italic_i = over¯ start_ARG 1 , italic_N - 1 end_ARG of the space grid are

Δhci=1h2(ci+12ci+ci1).subscriptΔsubscript𝑐𝑖1superscript2subscript𝑐𝑖12subscript𝑐𝑖subscript𝑐𝑖1\Delta_{h}c_{i}=\frac{1}{h^{2}}\left(c_{i+1}-2c_{i}+c_{i-1}\right).roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( italic_c start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT - 2 italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_c start_POSTSUBSCRIPT italic_i - 1 end_POSTSUBSCRIPT ) .

The finite difference equations

c0c1=0,cNcN1=0formulae-sequencesubscript𝑐0subscript𝑐10subscript𝑐𝑁subscript𝑐𝑁10c_{0}-c_{1}=0,\quad c_{N}-c_{N-1}=0italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0 , italic_c start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT - italic_c start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT = 0 (3.1)

approximating the homogeneous Neumann boundary conditions (2.2) are related to the boundary nodes i=0,N𝑖0𝑁i=0,Nitalic_i = 0 , italic_N. Then the resulting discrete Laplacian, with incorporated discretized boundary conditions, is denoted by ΔhsubscriptΔ\Delta_{h}roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT.

Explicit Euler (EE) scheme, the simplest discretization of (2.1), reads

1Mcin+1cinτ=Δhμin,μin=F(cin)ϵ2Δhcin.formulae-sequence1𝑀subscriptsuperscript𝑐𝑛1𝑖subscriptsuperscript𝑐𝑛𝑖τsubscriptΔsubscriptsuperscriptμ𝑛𝑖subscriptsuperscriptμ𝑛𝑖superscript𝐹subscriptsuperscript𝑐𝑛𝑖superscriptϵ2subscriptΔsubscriptsuperscript𝑐𝑛𝑖\displaystyle\frac{1}{M}\frac{c^{n+1}_{i}-c^{n}_{i}}{\tauup}=\Delta_{h}\muup^{% n}_{i},\quad\muup^{n}_{i}=F^{\prime}(c^{n}_{i})-\epsilonup^{2}\Delta_{h}c^{n}_% {i}.divide start_ARG 1 end_ARG start_ARG italic_M end_ARG divide start_ARG italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG roman_τ end_ARG = roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT roman_μ start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , roman_μ start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT .

The scheme is first order accurate in time and second order accurate in space, is stable for τh4similar-toτsuperscript4\tauup\sim h^{4}roman_τ ∼ italic_h start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT and is not energy stable.

Implicit Euler (IE) scheme reads

1Mcin+1cinτ=Δhμin+1,μin+1=F(cin+1)ϵ2Δhcin+1.formulae-sequence1𝑀subscriptsuperscript𝑐𝑛1𝑖subscriptsuperscript𝑐𝑛𝑖τsubscriptΔsubscriptsuperscriptμ𝑛1𝑖subscriptsuperscriptμ𝑛1𝑖superscript𝐹subscriptsuperscript𝑐𝑛1𝑖superscriptϵ2subscriptΔsubscriptsuperscript𝑐𝑛1𝑖\displaystyle\frac{1}{M}\frac{c^{n+1}_{i}-c^{n}_{i}}{\tauup}=\Delta_{h}\muup^{% n+1}_{i},\quad\muup^{n+1}_{i}=F^{\prime}(c^{n+1}_{i})-\epsilonup^{2}\Delta_{h}% c^{n+1}_{i}.divide start_ARG 1 end_ARG start_ARG italic_M end_ARG divide start_ARG italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG roman_τ end_ARG = roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT roman_μ start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , roman_μ start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT .

The scheme is first order accurate in time and second order accurate in space and is nonlinear (in the sense that a nonlinear equation has to be solved at each time step to determine cin+1subscriptsuperscript𝑐𝑛1𝑖c^{n+1}_{i}italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT). It is conditionally (i.e., for a sufficiently small ττ\tauuproman_τ) gradient stable and conditionally uniquely solvable, see Table 1.

Crank-Nicolson (CN) scheme can be written as

1Mcin+1cinτ=Δhμin+1/2,μin+1/2=12(μin+μin+1),formulae-sequence1𝑀subscriptsuperscript𝑐𝑛1𝑖subscriptsuperscript𝑐𝑛𝑖τsubscriptΔsubscriptsuperscriptμ𝑛12𝑖subscriptsuperscriptμ𝑛12𝑖12subscriptsuperscriptμ𝑛𝑖subscriptsuperscriptμ𝑛1𝑖\displaystyle\frac{1}{M}\frac{c^{n+1}_{i}-c^{n}_{i}}{\tauup}=\Delta_{h}\muup^{% n+1/2}_{i},\quad\muup^{n+1/2}_{i}=\frac{1}{2}(\muup^{n}_{i}+\muup^{n+1}_{i}),divide start_ARG 1 end_ARG start_ARG italic_M end_ARG divide start_ARG italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG roman_τ end_ARG = roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT roman_μ start_POSTSUPERSCRIPT italic_n + 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , roman_μ start_POSTSUPERSCRIPT italic_n + 1 / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( roman_μ start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + roman_μ start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ,

with μinsuperscriptsubscriptμ𝑖𝑛\muup_{i}^{n}roman_μ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT, μin+1superscriptsubscriptμ𝑖𝑛1\muup_{i}^{n+1}roman_μ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT defined above. The scheme is second order accurate in time and second order accurate in space and is nonlinear. It is conditionally gradient stable and conditionally uniquely solvable, see Table 1.

Semi-implicit Euler (SIE) reads

1Mcin+1cinτ=Δhμin+1,μin+1=F(cin)ϵ2Δhcin+1.formulae-sequence1𝑀subscriptsuperscript𝑐𝑛1𝑖subscriptsuperscript𝑐𝑛𝑖τsubscriptΔsubscriptsuperscriptμ𝑛1𝑖subscriptsuperscriptμ𝑛1𝑖superscript𝐹subscriptsuperscript𝑐𝑛𝑖superscriptϵ2subscriptΔsubscriptsuperscript𝑐𝑛1𝑖\displaystyle\frac{1}{M}\frac{c^{n+1}_{i}-c^{n}_{i}}{\tauup}=\Delta_{h}\muup^{% n+1}_{i},\quad\muup^{n+1}_{i}=F^{\prime}(c^{n}_{i})-\epsilonup^{2}\Delta_{h}c^% {n+1}_{i}.divide start_ARG 1 end_ARG start_ARG italic_M end_ARG divide start_ARG italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG roman_τ end_ARG = roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT roman_μ start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , roman_μ start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT .

The scheme is first order accurate in time and second order accurate in space and is linear. It is unconditionally uniquely solvable and is not energy stable, see Table 1.

3.2.  Eyre splitting

The time integration schemes considered up to now are not energy stable and/or not uniquely solvable for arbitrary large time step sizes [12, 13]. An efficient and general way to design time integration schemes, which are both uniquely solvable and energy stable for arbitrary time step sizes, is proposed in papers of David J. Eyre [14, 15]. Since we essentially employ this approach to construct new schemes, we discuss it in more details here.

The essence of the Eyre approach is as follows. The system free energy density ψψ\psiuproman_ψ, c.f. (2.4), is not a convex function. This is because the dividing part of the free energy is a double-well potential. Assume that ψψ\psiuproman_ψ can be split as

ψ=ψc+ψe,ψsubscriptψcsubscriptψe\psiup=\psiup_{\text{c}}+\psiup_{\text{e}},roman_ψ = roman_ψ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT + roman_ψ start_POSTSUBSCRIPT e end_POSTSUBSCRIPT , (3.2)

where ψcsubscriptψc\psiup_{\text{c}}roman_ψ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT and ψesubscriptψe-\psiup_{\text{e}}- roman_ψ start_POSTSUBSCRIPT e end_POSTSUBSCRIPT are convex functions. The subindices ‘‘c’’ (meaning ‘‘contraction’’) and ‘‘e’’ (meaning ‘‘expansion’’) reflect their physical meaning. Here the convexity of a function is understood as a positive semi-definiteness of its Hessian, see [13], i.e., a function f(x)𝑓𝑥f(x)italic_f ( italic_x ) is convex provided that f′′(x)d2f/dx20superscript𝑓′′𝑥superscript𝑑2𝑓𝑑superscript𝑥20f^{\prime\prime}(x)\equiv d^{2}f/dx^{2}\geqslant 0italic_f start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_x ) ≡ italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_f / italic_d italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⩾ 0. Then, the semi-discrete (discrete in time) approximations

1Mcn+1cnτ=Δ[μc(cn+1)+μe(cn)],1𝑀superscript𝑐𝑛1superscript𝑐𝑛τΔdelimited-[]subscriptμcsuperscript𝑐𝑛1subscriptμesuperscript𝑐𝑛\displaystyle\frac{1}{M}\frac{c^{n+1}-c^{n}}{\tauup}=\Delta\left[\muup_{\text{% c}}(c^{n+1})+\muup_{\text{e}}(c^{n})\right],divide start_ARG 1 end_ARG start_ARG italic_M end_ARG divide start_ARG italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT - italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG roman_τ end_ARG = roman_Δ [ roman_μ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT ( italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) + roman_μ start_POSTSUBSCRIPT e end_POSTSUBSCRIPT ( italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ] , (3.3)

with

μc,e=δψc,eδc,μ=μc+μe,formulae-sequencesubscriptμc,eδsubscriptψc,eδ𝑐μsubscriptμcsubscriptμe\muup_{\text{c,e}}=\frac{\deltaup\psiup_{\text{c,e}}}{\deltaup c},\quad\muup=% \muup_{\text{c}}+\muup_{\text{e}},roman_μ start_POSTSUBSCRIPT c,e end_POSTSUBSCRIPT = divide start_ARG roman_δ roman_ψ start_POSTSUBSCRIPT c,e end_POSTSUBSCRIPT end_ARG start_ARG roman_δ italic_c end_ARG , roman_μ = roman_μ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT + roman_μ start_POSTSUBSCRIPT e end_POSTSUBSCRIPT ,

are gradient stable and

ψ(cn+1)ψ(cn)ψsuperscript𝑐𝑛1ψsuperscript𝑐𝑛\psiup(c^{n+1})\leqslant\psiup(c^{n})roman_ψ ( italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ⩽ roman_ψ ( italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT )

holds for any time step size τ>0τ0\tauup>0roman_τ > 0. In addition, the convexity of the energy partψcsubscriptψc\psiup_{\text{c}}roman_ψ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT corresponding to the implicitly treated term, ensures the unique solvability of (3.3).

Consider the specific form of the energy functional (2.4). Taking into account that the gradient term in (2.4) belongs to the convex part of the energy, we can take in (3.2)

ψc=12|c|2+Fc,ψe=Fe,Fc+Fe=F.formulae-sequencesubscriptψc12superscript𝑐2subscript𝐹cformulae-sequencesubscriptψesubscript𝐹esubscript𝐹csubscript𝐹e𝐹\psiup_{\text{c}}=\frac{1}{2}|\nabla c|^{2}+F_{\text{c}},\quad\psiup_{\text{e}% }=F_{\text{e}},\quad F_{\text{c}}+F_{\text{e}}=F.roman_ψ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG | ∇ italic_c | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_F start_POSTSUBSCRIPT c end_POSTSUBSCRIPT , roman_ψ start_POSTSUBSCRIPT e end_POSTSUBSCRIPT = italic_F start_POSTSUBSCRIPT e end_POSTSUBSCRIPT , italic_F start_POSTSUBSCRIPT c end_POSTSUBSCRIPT + italic_F start_POSTSUBSCRIPT e end_POSTSUBSCRIPT = italic_F .

Thus, constructing a suitable splitting of the system energy is reduced to splitting the dividing part of the energy F𝐹Fitalic_F in such a way that

F=Fc+Fe,Fc′′0,Fe′′0.formulae-sequence𝐹subscript𝐹csubscript𝐹eformulae-sequencesubscriptsuperscript𝐹′′c0subscriptsuperscript𝐹′′e0F=F_{\text{c}}+F_{\text{e}},\quad F^{\prime\prime}_{\text{c}}\geqslant 0,\quad% -F^{\prime\prime}_{\text{e}}\geqslant 0.italic_F = italic_F start_POSTSUBSCRIPT c end_POSTSUBSCRIPT + italic_F start_POSTSUBSCRIPT e end_POSTSUBSCRIPT , italic_F start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT c end_POSTSUBSCRIPT ⩾ 0 , - italic_F start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT e end_POSTSUBSCRIPT ⩾ 0 . (3.4)

Then, the chemical potential takes form

μ=μc+μe,μc=ϵ2Δc+Fc,μe=Fe,formulae-sequenceμsubscriptμcsubscriptμeformulae-sequencesubscriptμcsuperscriptϵ2Δ𝑐subscriptsuperscript𝐹csubscriptμesubscriptsuperscript𝐹e\muup=\muup_{\text{c}}+\muup_{\text{e}},\quad\muup_{\text{c}}=-\epsilonup^{2}% \Delta c+F^{\prime}_{\text{c}},\quad\muup_{\text{e}}=F^{\prime}_{\text{e}},roman_μ = roman_μ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT + roman_μ start_POSTSUBSCRIPT e end_POSTSUBSCRIPT , roman_μ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = - roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ italic_c + italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT c end_POSTSUBSCRIPT , roman_μ start_POSTSUBSCRIPT e end_POSTSUBSCRIPT = italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT e end_POSTSUBSCRIPT ,

whereas the semi-discrete system (3.3) can be written as

1Mcn+1cnτ=ϵ2Δ2cn+1+ΔFc(cn+1)+ΔFe(cn).1𝑀superscript𝑐𝑛1superscript𝑐𝑛τsuperscriptϵ2superscriptΔ2superscript𝑐𝑛1Δsubscriptsuperscript𝐹csuperscript𝑐𝑛1Δsubscriptsuperscript𝐹esuperscript𝑐𝑛\frac{1}{M}\frac{c^{n+1}-c^{n}}{\tauup}=-\epsilonup^{2}\Delta^{2}c^{n+1}+% \Delta F^{\prime}_{\text{c}}(c^{n+1})+\Delta F^{\prime}_{\text{e}}(c^{n}).divide start_ARG 1 end_ARG start_ARG italic_M end_ARG divide start_ARG italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT - italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG roman_τ end_ARG = - roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT + roman_Δ italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT c end_POSTSUBSCRIPT ( italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) + roman_Δ italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT e end_POSTSUBSCRIPT ( italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) . (3.5)

Splitting of the form (3.4) can be constructed in a number of ways. One of them is to write the dividing part F𝐹Fitalic_F of the free energy as a sum of two terms according to (3.2). Note that the gradient part of the free energy (2.5) should be handled together with the convex part ψcsubscriptψc\psiup_{\text{c}}roman_ψ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT. Another way is to choose a regularizing function Frsubscript𝐹rF_{\text{r}}italic_F start_POSTSUBSCRIPT r end_POSTSUBSCRIPT such that

Fc=F+Fr,Fe=FFr.formulae-sequencesubscript𝐹c𝐹subscript𝐹rsubscript𝐹e𝐹subscript𝐹rF_{\text{c}}=F+F_{\text{r}},\quad F_{\text{e}}=F-F_{\text{r}}.italic_F start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = italic_F + italic_F start_POSTSUBSCRIPT r end_POSTSUBSCRIPT , italic_F start_POSTSUBSCRIPT e end_POSTSUBSCRIPT = italic_F - italic_F start_POSTSUBSCRIPT r end_POSTSUBSCRIPT .

Although both ways are essentially quite similar, it is convenient to distinguish them due to various possible generalizations.

Consider the specific form (2.5) of F𝐹Fitalic_F used in this work. We have

F(c)=c2(1c2),F(c)=2c(2c23c+1),F′′(c)=2(6c26c+1).formulae-sequence𝐹𝑐superscript𝑐21superscript𝑐2formulae-sequencesuperscript𝐹𝑐2𝑐2superscript𝑐23𝑐1superscript𝐹′′𝑐26superscript𝑐26𝑐1F(c)=c^{2}(1-c^{2}),\quad F^{\prime}(c)=2c(2c^{2}-3c+1),\quad F^{\prime\prime}% (c)=2(6c^{2}-6c+1).italic_F ( italic_c ) = italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) , italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_c ) = 2 italic_c ( 2 italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 3 italic_c + 1 ) , italic_F start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_c ) = 2 ( 6 italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 6 italic_c + 1 ) .

Equation F′′(c)=0superscript𝐹′′𝑐0F^{\prime\prime}(c)=0italic_F start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_c ) = 0 has roots c1,2=(3±3)/6subscript𝑐12plus-or-minus336c_{1,2}=\left(3\pm\sqrt{3}\right)/6italic_c start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT = ( 3 ± square-root start_ARG 3 end_ARG ) / 6, and, hence, F′′(c)<0superscript𝐹′′𝑐0F^{\prime\prime}(c)<0italic_F start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_c ) < 0 for (33)/6<c<(3+3)/6336𝑐336\left(3-\sqrt{3}\right)/6<c<\left(3+\sqrt{3}\right)/6( 3 - square-root start_ARG 3 end_ARG ) / 6 < italic_c < ( 3 + square-root start_ARG 3 end_ARG ) / 6 ; F′′(c)0superscript𝐹′′𝑐0F^{\prime\prime}(c)\geqslant 0italic_F start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_c ) ⩾ 0 for c(33)/6𝑐336c\leqslant\left(3-\sqrt{3}\right)/6italic_c ⩽ ( 3 - square-root start_ARG 3 end_ARG ) / 6 and c(3+3)/6𝑐336c\geqslant\left(3+\sqrt{3}\right)/6italic_c ⩾ ( 3 + square-root start_ARG 3 end_ARG ) / 6. Function F′′(c)superscript𝐹′′𝑐F^{\prime\prime}(c)italic_F start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_c ) attains its minimum, equal to 11-1- 1, at c=1/2𝑐12c=1/2italic_c = 1 / 2. The maxima of F′′(c)superscript𝐹′′𝑐F^{\prime\prime}(c)italic_F start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_c ) for c[0,1]𝑐01c\in[0,1]italic_c ∈ [ 0 , 1 ] are attained at c=0𝑐0c=0italic_c = 0 and c=1𝑐1c=1italic_c = 1 and equal 2222. Thus, to ensure that Fcsubscript𝐹cF_{\text{c}}italic_F start_POSTSUBSCRIPT c end_POSTSUBSCRIPT and Fesubscript𝐹e-F_{\text{e}}- italic_F start_POSTSUBSCRIPT e end_POSTSUBSCRIPT are convex, it suffices to set

Fc=F(c)+12c2,Fe=12c2.formulae-sequencesubscript𝐹c𝐹𝑐12superscript𝑐2subscript𝐹e12superscript𝑐2F_{\text{c}}=F(c)+\frac{1}{2}c^{2},\quad F_{\text{e}}=-\frac{1}{2}c^{2}.italic_F start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = italic_F ( italic_c ) + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , italic_F start_POSTSUBSCRIPT e end_POSTSUBSCRIPT = - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT .

This leads to non-linearly stabilized splitting scheme (NLSS),

1Mcin+1cinτ=Δhμn+1,μin+1=[F(cin+1)+cin+1]cinϵ2Δhcin+1.formulae-sequence1𝑀subscriptsuperscript𝑐𝑛1𝑖subscriptsuperscript𝑐𝑛𝑖τsubscriptΔsuperscriptμ𝑛1subscriptsuperscriptμ𝑛1𝑖delimited-[]superscript𝐹subscriptsuperscript𝑐𝑛1𝑖subscriptsuperscript𝑐𝑛1𝑖subscriptsuperscript𝑐𝑛𝑖superscriptϵ2subscriptΔsubscriptsuperscript𝑐𝑛1𝑖\displaystyle\frac{1}{M}\frac{c^{n+1}_{i}-c^{n}_{i}}{\tauup}=\Delta_{h}\muup^{% n+1},\quad\muup^{n+1}_{i}=\left[F^{\prime}(c^{n+1}_{i})+c^{n+1}_{i}\right]-c^{% n}_{i}-\epsilonup^{2}\Delta_{h}c^{n+1}_{i}.divide start_ARG 1 end_ARG start_ARG italic_M end_ARG divide start_ARG italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG roman_τ end_ARG = roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT roman_μ start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT , roman_μ start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = [ italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) + italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] - italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT . (3.6)

The scheme can also be written as

1Mcin+1cinτ=(Δhϵ2Δh2)cin+1+ΔhF(cin+1)Δcin.1𝑀subscriptsuperscript𝑐𝑛1𝑖subscriptsuperscript𝑐𝑛𝑖τsubscriptΔsuperscriptϵ2superscriptsubscriptΔ2subscriptsuperscript𝑐𝑛1𝑖subscriptΔsuperscript𝐹subscriptsuperscript𝑐𝑛1𝑖Δsubscriptsuperscript𝑐𝑛𝑖\frac{1}{M}\frac{c^{n+1}_{i}-c^{n}_{i}}{\tauup}=\left(\Delta_{h}-\epsilonup^{2% }\Delta_{h}^{2}\right)c^{n+1}_{i}+\Delta_{h}F^{\prime}(c^{n+1}_{i})-\Delta c^{% n}_{i}.divide start_ARG 1 end_ARG start_ARG italic_M end_ARG divide start_ARG italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG roman_τ end_ARG = ( roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT - roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - roman_Δ italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT .

It is first order accurate in time, second order in space, implicit, nonlinear in cn+1superscript𝑐𝑛1c^{n+1}italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT and energy stable for arbitrary τ>0τ0\tauup>0roman_τ > 0.

Another option is to set

Fc=c2,Fe=c2+F(c),formulae-sequencesubscript𝐹csuperscript𝑐2subscript𝐹esuperscript𝑐2𝐹𝑐F_{\text{c}}=c^{2},\quad F_{\text{e}}=-c^{2}+F(c),italic_F start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , italic_F start_POSTSUBSCRIPT e end_POSTSUBSCRIPT = - italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_F ( italic_c ) ,

which leads to linearly stabilized splitting scheme (LSS),

1Mcin+1cinτ=Δhμin+1,μin+1=2cin+1+[F(cin)2cin]ϵ2Δhcin+1.formulae-sequence1𝑀subscriptsuperscript𝑐𝑛1𝑖subscriptsuperscript𝑐𝑛𝑖τsubscriptΔsubscriptsuperscriptμ𝑛1𝑖subscriptsuperscriptμ𝑛1𝑖2subscriptsuperscript𝑐𝑛1𝑖delimited-[]superscript𝐹subscriptsuperscript𝑐𝑛𝑖2subscriptsuperscript𝑐𝑛𝑖superscriptϵ2subscriptΔsubscriptsuperscript𝑐𝑛1𝑖\displaystyle\frac{1}{M}\frac{c^{n+1}_{i}-c^{n}_{i}}{\tauup}=\Delta_{h}\muup^{% n+1}_{i},\quad\muup^{n+1}_{i}=2c^{n+1}_{i}+\left[F^{\prime}(c^{n}_{i})-2c^{n}_% {i}\right]-\epsilonup^{2}\Delta_{h}c^{n+1}_{i}.divide start_ARG 1 end_ARG start_ARG italic_M end_ARG divide start_ARG italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG roman_τ end_ARG = roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT roman_μ start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , roman_μ start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 2 italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + [ italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - 2 italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ] - roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT .

It may be convenient to write it as

1Mcin+1cinτ=(2Δhϵ2Δh2)cin+1+ΔhF(cin)2Δcin.1𝑀subscriptsuperscript𝑐𝑛1𝑖subscriptsuperscript𝑐𝑛𝑖τ2subscriptΔsuperscriptϵ2superscriptsubscriptΔ2subscriptsuperscript𝑐𝑛1𝑖subscriptΔsuperscript𝐹subscriptsuperscript𝑐𝑛𝑖2Δsubscriptsuperscript𝑐𝑛𝑖\frac{1}{M}\frac{c^{n+1}_{i}-c^{n}_{i}}{\tauup}=\left(2\Delta_{h}-\epsilonup^{% 2}\Delta_{h}^{2}\right)c^{n+1}_{i}+\Delta_{h}F^{\prime}(c^{n}_{i})-2\Delta c^{% n}_{i}.divide start_ARG 1 end_ARG start_ARG italic_M end_ARG divide start_ARG italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG roman_τ end_ARG = ( 2 roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT - roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - 2 roman_Δ italic_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT .

The scheme is first order accurate in time, second order in space, implicit and, provided that splitting parameters are chosen properly [40], energy stable for arbitrary τ>0τ0\tauup>0roman_τ > 0. Unlike the NLSS scheme (3.6), it is linear with respect to cn+1superscript𝑐𝑛1c^{n+1}italic_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT. A three-parameter family of linearly stabilized schemes, including the LSS scheme, is analyzed in [13].

Properties of considered time integration schemes are summarized in Table 1.

Table 1: Properties of time integration schemes according to [12, 13, 14, 15, 40] ( provided parameters are chosen properly, see [40]).
Scheme Linearity Gradient stability Solvability
Explicit Euler (EE) Yes No Yes
Implicit Euler (IE) No Conditional, τ14h2τ14superscript2\tauup\leqslant\frac{1}{4}h^{2}roman_τ ⩽ divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT τ118h2τ118superscript2\tauup\leqslant\frac{1}{18}h^{2}roman_τ ⩽ divide start_ARG 1 end_ARG start_ARG 18 end_ARG italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
Crank-Nicolson (CN) No Conditional τ19h2τ19superscript2\tauup\leqslant\frac{1}{9}h^{2}roman_τ ⩽ divide start_ARG 1 end_ARG start_ARG 9 end_ARG italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
Semi-implicit Euler (SIE) Yes No Yes
Nonlinearly stabilized splitting (NLSS) No Unconditional Yes
Linearly stabilized splitting (LSS) Yes Unconditional Yes

For a fully discrete scheme (i.e., when discretization in time and in space is applied), its gradient stability implies that the free energy functional (2.4) is approximated in a certain way. In this work such an approximation, determined by the space disretization, reads

ψh(𝐜)=hi=1NF(ci)+12ϵ2hi=1N(ci+1ci)2h2.subscriptψ𝐜superscriptsubscript𝑖1𝑁superscript𝐹subscript𝑐𝑖12superscriptϵ2superscriptsubscript𝑖1𝑁superscriptsubscript𝑐𝑖1subscript𝑐𝑖2superscript2\psiup_{h}(\mathbf{c})=h\sum\limits_{i=1}^{N}F^{\prime}(c_{i})+\frac{1}{2}% \epsilonup^{2}h\sum\limits_{i=1}^{N}\frac{(c_{i+1}-c_{i})^{2}}{h^{2}}.roman_ψ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( bold_c ) = italic_h ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) + divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_h ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT divide start_ARG ( italic_c start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT - italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (3.7)

Let a N×N𝑁𝑁N\times Nitalic_N × italic_N matrix 𝐀𝐀\mathbf{A}bold_A be the matrix of the discussed above discretization ΔhsubscriptΔ-\Delta_{h}- roman_Δ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT of the operator ΔΔ-\Delta- roman_Δ, with incorporated homogeneous Neumann boundary conditions (3.1). The space discretization reduces the initial boundary-value problem (2.1)–(2.3) to an initial value problem

1M𝐜=𝐀(F(𝐜)+ϵ2𝐀𝐜),𝐜(0)=𝐜0,formulae-sequence1𝑀superscript𝐜𝐀superscript𝐹𝐜superscriptϵ2𝐀𝐜𝐜0superscript𝐜0\frac{1}{M}\mathbf{c}^{\prime}=-\mathbf{A}\left(F^{\prime}(\mathbf{c})+% \epsilonup^{2}\mathbf{A}\mathbf{c}\right),\quad\mathbf{c}(0)=\mathbf{c}^{0},divide start_ARG 1 end_ARG start_ARG italic_M end_ARG bold_c start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = - bold_A ( italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c ) + roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_Ac ) , bold_c ( 0 ) = bold_c start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT , (3.8)

where the entries of the vector function 𝐜(t):N:𝐜𝑡superscript𝑁\mathbf{c}(t):\mathbb{R}\rightarrow\mathbb{R}^{N}bold_c ( italic_t ) : blackboard_R → blackboard_R start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT are the sought after solution values on the space grid and the vector 𝐜0superscript𝐜0\mathbf{c}^{0}bold_c start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT contains the grid values of the given function c0(x)superscript𝑐0𝑥c^{0}(x)italic_c start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ( italic_x ).

Let 𝐜nsuperscript𝐜𝑛\mathbf{c}^{n}bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT be an N𝑁Nitalic_N vector containing the numerical solution values for time level n𝑛nitalic_n. Then, the time integration schemes discussed above can be written in a compact form as follows.

  • Explicit Euler (EE):

    𝐜n+1𝐜nτ=𝐀(F(𝐜n)+ϵ2𝐀𝐜n);superscript𝐜𝑛1superscript𝐜𝑛τ𝐀superscript𝐹superscript𝐜𝑛superscriptϵ2superscript𝐀𝐜𝑛\frac{\mathbf{c}^{n+1}-\mathbf{c}^{n}}{\tauup}=-\mathbf{A}\left(F^{\prime}(% \mathbf{c}^{n})+\epsilonup^{2}\mathbf{A}\mathbf{c}^{n}\right);divide start_ARG bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT - bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG roman_τ end_ARG = - bold_A ( italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) + roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_Ac start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ; (3.9)
  • implicit Euler (IE):

    𝐜n+1𝐜nτ=𝐀(F(𝐜n+1)+ϵ2𝐀𝐜n+1);superscript𝐜𝑛1superscript𝐜𝑛τ𝐀superscript𝐹superscript𝐜𝑛1superscriptϵ2superscript𝐀𝐜𝑛1\frac{\mathbf{c}^{n+1}-\mathbf{c}^{n}}{\tauup}=-\mathbf{A}\left(F^{\prime}(% \mathbf{c}^{n+1})+\epsilonup^{2}\mathbf{A}\mathbf{c}^{n+1}\right);divide start_ARG bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT - bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG roman_τ end_ARG = - bold_A ( italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) + roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_Ac start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ; (3.10)
  • Crank-Nicolson (CN):

    𝐜n+1𝐜nτ=12𝐀(F(𝐜n+1)+ϵ2𝐀𝐜n+1+F(𝐜n)+ϵ2𝐀𝐜n);superscript𝐜𝑛1superscript𝐜𝑛τ12𝐀superscript𝐹superscript𝐜𝑛1superscriptϵ2superscript𝐀𝐜𝑛1superscript𝐹superscript𝐜𝑛superscriptϵ2superscript𝐀𝐜𝑛\frac{\mathbf{c}^{n+1}-\mathbf{c}^{n}}{\tauup}=-\frac{1}{2}\mathbf{A}\left(F^{% \prime}(\mathbf{c}^{n+1})+\epsilonup^{2}\mathbf{A}\mathbf{c}^{n+1}+F^{\prime}(% \mathbf{c}^{n})+\epsilonup^{2}\mathbf{A}\mathbf{c}^{n}\right);divide start_ARG bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT - bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG roman_τ end_ARG = - divide start_ARG 1 end_ARG start_ARG 2 end_ARG bold_A ( italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) + roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_Ac start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT + italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) + roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_Ac start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ; (3.11)
  • Semi-implicit Euler (SIE):

    𝐜n+1𝐜nτ=𝐀(F(𝐜n)+ϵ2𝐀𝐜n+1);superscript𝐜𝑛1superscript𝐜𝑛τ𝐀superscript𝐹superscript𝐜𝑛superscriptϵ2superscript𝐀𝐜𝑛1\frac{\mathbf{c}^{n+1}-\mathbf{c}^{n}}{\tauup}=-\mathbf{A}\left(F^{\prime}(% \mathbf{c}^{n})+\epsilonup^{2}\mathbf{A}\mathbf{c}^{n+1}\right);divide start_ARG bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT - bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG roman_τ end_ARG = - bold_A ( italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) + roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_Ac start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ; (3.12)
  • nonlinearly stabilized splitting (NLSS)

    𝐜n+1𝐜nτ=𝐀(F(𝐜n+1)+𝐜n+1𝐜n+ϵ2𝐀𝐜n+1);superscript𝐜𝑛1superscript𝐜𝑛τ𝐀superscript𝐹superscript𝐜𝑛1superscript𝐜𝑛1superscript𝐜𝑛superscriptϵ2superscript𝐀𝐜𝑛1\frac{\mathbf{c}^{n+1}-\mathbf{c}^{n}}{\tauup}=-\mathbf{A}\left(F^{\prime}(% \mathbf{c}^{n+1})+\mathbf{c}^{n+1}-\mathbf{c}^{n}+\epsilonup^{2}\mathbf{A}% \mathbf{c}^{n+1}\right);divide start_ARG bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT - bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG roman_τ end_ARG = - bold_A ( italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) + bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT - bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT + roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_Ac start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ; (3.13)
  • linearly stabilized splitting (LSS)

    𝐜n+1𝐜nτ=𝐀(F(𝐜n)2𝐜n+2𝐜n+1+ϵ2𝐀𝐜n+1).superscript𝐜𝑛1superscript𝐜𝑛τ𝐀superscript𝐹superscript𝐜𝑛2superscript𝐜𝑛2superscript𝐜𝑛1superscriptϵ2superscript𝐀𝐜𝑛1\frac{\mathbf{c}^{n+1}-\mathbf{c}^{n}}{\tauup}=-\mathbf{A}\left(F^{\prime}(% \mathbf{c}^{n})-2\mathbf{c}^{n}+2\mathbf{c}^{n+1}+\epsilonup^{2}\mathbf{A}% \mathbf{c}^{n+1}\right).divide start_ARG bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT - bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG roman_τ end_ARG = - bold_A ( italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) - 2 bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT + 2 bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT + roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_Ac start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) . (3.14)

We now briefly discuss computational complexity of these schemes. In implicit nonlinear schemes (i.e., IE, CN, NLSS), each time step a nonlinear system of equations has to be solved, e.g., by a Newton iteration. In turn, each Newton iteration involves solving the Jacobian linear system. Implicit linear schemes SIE and LSS require solving a single linear system per time step. Explicit schemes involve neither nonlinear nor linear system solution and have lowest costs per time step. Although implicit time integration schemes allow to use significantly larger time step sizes than explicit schemes, solving large linear systems, within Newton iterations or independently, may become a computationally expensive task. This is because efficiency in sparse direct solvers, as well as in preconditioned iterative solvers, is difficult to retain on modern high-performance hybrid CPU/GPU platforms.

The aim of this work is to develop an efficient numerical algorithm which, on one hand, is gradient stable and allows large time step sizes and, on the other hand, can be efficiently implemented on modern high-performance hybrid platforms. Our approach is based on the combination of the Eyre splitting, in particular the LSS scheme, with the local iteration modified (LIM) scheme. Then, the Eyre splitting allows to use large time step sizes while keeping gradient stability, whereas the LIM mechanism provides a computationally efficient alternative to solving linear systems in implicit schemes.

3.3.  Local iteration modified (LIM) scheme

The local iteration modified (LIM) scheme can be viewed as a special explicit scheme where a Chebyshev polynomial of a sufficiently high order p𝑝pitalic_p is employed to ensure stability and monotonicity for a given time step size τ>0τ0\tauup>0roman_τ > 0. We now describe a LIM variant based on the linearly stabilized splitting (LSS) scheme (3.14), which we write as

𝐜n+1=[𝐈+τ𝐀^]1𝐟^n,𝐀^=𝐀(2𝐈+ϵ2𝐀),𝐟^n=𝐜n+τ𝐀(2𝐜nF(𝐜n)),formulae-sequencesuperscript𝐜𝑛1superscriptdelimited-[]𝐈τ^𝐀1superscript^𝐟𝑛formulae-sequence^𝐀𝐀2𝐈superscriptϵ2𝐀superscript^𝐟𝑛superscript𝐜𝑛τ𝐀2superscript𝐜𝑛superscript𝐹superscript𝐜𝑛\mathbf{c}^{n+1}=\left[\mathbf{I}+\tauup\widehat{\mathbf{A}}\right]^{-1}% \widehat{\mathbf{f}}^{n},\quad\widehat{\mathbf{A}}=\mathbf{A}(2\mathbf{I}+% \epsilonup^{2}\mathbf{A}),\quad\widehat{\mathbf{f}}^{n}=\mathbf{c}^{n}+\tauup% \mathbf{A}(2\mathbf{c}^{n}-F^{\prime}(\mathbf{c}^{n})),bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT = [ bold_I + roman_τ over^ start_ARG bold_A end_ARG ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT over^ start_ARG bold_f end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT , over^ start_ARG bold_A end_ARG = bold_A ( 2 bold_I + roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_A ) , over^ start_ARG bold_f end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT + roman_τ bold_A ( 2 bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ) , (3.15)

where 𝐈𝐈\mathbf{I}bold_I is the N×N𝑁𝑁N\times Nitalic_N × italic_N identity matrix. We emphasize that the inverse matrix here is, of course, never computed, rather a linear system with this matrix is solved. Within the LIM approach the inverse matrix is replaced by a specially chosen Chebyshev polynomial. Let λsubscriptλ\lambdaup_{\infty}roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT be an upper bound for the largest eigenvalue of 𝐀^^𝐀\widehat{\mathbf{A}}over^ start_ARG bold_A end_ARG (in practice one can set λ:=𝐀^1=maxji|a^ij|assignsubscriptλsubscriptnorm^𝐀1subscript𝑗subscript𝑖subscript^𝑎𝑖𝑗\lambdaup_{\infty}:=\|\widehat{\mathbf{A}}\|_{1}=\max_{j}\sum_{i}|\widehat{a}_% {ij}|roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT := ∥ over^ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_max start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | over^ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT |). The polynomial order p𝑝pitalic_p for which stability takes place is chosen as

p=π4τλ+1,𝑝π4τsubscriptλ1p=\left\lceil\frac{\piup}{4}\sqrt{\tauup\lambdaup_{\infty}+1}\,\right\rceil,italic_p = ⌈ divide start_ARG roman_π end_ARG start_ARG 4 end_ARG square-root start_ARG roman_τ roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT + 1 end_ARG ⌉ , (3.16)

where x𝑥\lceil x\rceil⌈ italic_x ⌉, for x𝑥x\in\mathbb{R}italic_x ∈ blackboard_R, denotes the smallest integer greater than or equal to x𝑥xitalic_x. Let the Chebyshev polynomial roots

{βm,m=1,,p}={cosπ2i12p,i=1,,p}\{\betaup_{m},\;m=1,\ldots,p\}=\left\{\cos\piup\frac{2i-1}{2p},\;i=1,\dots,p\right\}{ roman_β start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT , italic_m = 1 , … , italic_p } = { roman_cos roman_π divide start_ARG 2 italic_i - 1 end_ARG start_ARG 2 italic_p end_ARG , italic_i = 1 , … , italic_p }

be ordered in such a way that no numerical instability occurs for arbitrarily large p𝑝pitalic_p and β1=cos(π/2p)subscriptβ1π2𝑝\betaup_{1}=\cos({\piup}/{2p})roman_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_cos ( roman_π / 2 italic_p ) is the first root. We set z1=β1subscript𝑧1subscriptβ1z_{1}=\betaup_{1}italic_z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and define Chebyshev polynomial parameters

am=λ1+z1(z1βm),m=1,,p.formulae-sequencesubscript𝑎𝑚subscriptλ1subscript𝑧1subscript𝑧1subscriptβ𝑚𝑚1𝑝a_{m}=\frac{\lambdaup_{\infty}}{1+z_{1}}(z_{1}-\betaup_{m}),\quad m=1,\dots,p.italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = divide start_ARG roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG start_ARG 1 + italic_z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG ( italic_z start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - roman_β start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) , italic_m = 1 , … , italic_p .

Denoting 𝐲(0)=𝐜nsuperscript𝐲0superscript𝐜𝑛\mathbf{y}^{(0)}=\mathbf{c}^{n}bold_y start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT = bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT, we obtain the LIM solution 𝐜n+1superscript𝐜𝑛1\mathbf{c}^{n+1}bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT on the next time level by carrying out 2p12𝑝12p-12 italic_p - 1 Chebyshev iterations as

𝐲(m)superscript𝐲𝑚\displaystyle\mathbf{y}^{(m)}bold_y start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT =11+τam(𝐜n+τam𝐲(m1)+τ(𝐟^n𝐀^𝐲(m1))),m=1,,p,formulae-sequenceabsent11τsubscript𝑎𝑚superscript𝐜𝑛τsubscript𝑎𝑚superscript𝐲𝑚1τsuperscript^𝐟𝑛^𝐀superscript𝐲𝑚1𝑚1𝑝\displaystyle=\frac{1}{1+\tauup a_{m}}\left(\mathbf{c}^{n}+\tauup a_{m}\mathbf% {y}^{(m-1)}+\tauup(\widehat{\mathbf{f}}^{n}-\widehat{\mathbf{A}}\mathbf{y}^{(m% -1)})\right),\quad m=1,\dots,p,= divide start_ARG 1 end_ARG start_ARG 1 + roman_τ italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT + roman_τ italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT bold_y start_POSTSUPERSCRIPT ( italic_m - 1 ) end_POSTSUPERSCRIPT + roman_τ ( over^ start_ARG bold_f end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - over^ start_ARG bold_A end_ARG bold_y start_POSTSUPERSCRIPT ( italic_m - 1 ) end_POSTSUPERSCRIPT ) ) , italic_m = 1 , … , italic_p , (3.17)
𝐲(p+m1)superscript𝐲𝑝𝑚1\displaystyle\mathbf{y}^{(p+m-1)}bold_y start_POSTSUPERSCRIPT ( italic_p + italic_m - 1 ) end_POSTSUPERSCRIPT =11+τam(𝐜n+τam𝐲(m1)+τ(𝐟^n𝐀^𝐲(m1))),m=2,,p,formulae-sequenceabsent11τsubscript𝑎𝑚superscript𝐜𝑛τsubscript𝑎𝑚superscript𝐲𝑚1τsuperscript^𝐟𝑛^𝐀superscript𝐲𝑚1𝑚2𝑝\displaystyle=\frac{1}{1+\tauup a_{m}}\left(\mathbf{c}^{n}+\tauup a_{m}\mathbf% {y}^{(m-1)}+\tauup(\widehat{\mathbf{f}}^{n}-\widehat{\mathbf{A}}\mathbf{y}^{(m% -1)})\right),\quad m=2,\dots,p,= divide start_ARG 1 end_ARG start_ARG 1 + roman_τ italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT + roman_τ italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT bold_y start_POSTSUPERSCRIPT ( italic_m - 1 ) end_POSTSUPERSCRIPT + roman_τ ( over^ start_ARG bold_f end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - over^ start_ARG bold_A end_ARG bold_y start_POSTSUPERSCRIPT ( italic_m - 1 ) end_POSTSUPERSCRIPT ) ) , italic_m = 2 , … , italic_p ,
𝐜n+1superscript𝐜𝑛1\displaystyle\mathbf{c}^{n+1}bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT =𝐲(2p1).absentsuperscript𝐲2𝑝1\displaystyle=\mathbf{y}^{(2p-1)}.= bold_y start_POSTSUPERSCRIPT ( 2 italic_p - 1 ) end_POSTSUPERSCRIPT .

Note that a1=0subscript𝑎10a_{1}=0italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0 and it is easy to check that the first iteration solution 𝐲(1)superscript𝐲1\mathbf{y}^{(1)}bold_y start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT is the explicit Euler solution on the next time level. In this sense, the Chebyshev iterations actually start at the second step by computing 𝐲(2)superscript𝐲2\mathbf{y}^{(2)}bold_y start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT. If only the first sweep of iterations is carried out in (3.17), with 𝐲(1)superscript𝐲1\mathbf{y}^{(1)}bold_y start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT, …, 𝐲(p)superscript𝐲𝑝\mathbf{y}^{(p)}bold_y start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT computed, then setting 𝐜n+1=𝐲(p)superscript𝐜𝑛1superscript𝐲𝑝\mathbf{c}^{n+1}=\mathbf{y}^{(p)}bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT = bold_y start_POSTSUPERSCRIPT ( italic_p ) end_POSTSUPERSCRIPT we get a time integration scheme which is known as the regular (non-modified) local iteration scheme (the LI scheme). As already mentioned, in the regular LI scheme stability of the numerical solution is guaranteed for parabolic problems but the solution nonnegativity is not. As one can see, in the LIM scheme (3.17) the second Chebyshev iteration sweep, with 𝐲(p+1)superscript𝐲𝑝1\mathbf{y}^{(p+1)}bold_y start_POSTSUPERSCRIPT ( italic_p + 1 ) end_POSTSUPERSCRIPT, …, 𝐲(2p1)superscript𝐲2𝑝1\mathbf{y}^{(2p-1)}bold_y start_POSTSUPERSCRIPT ( 2 italic_p - 1 ) end_POSTSUPERSCRIPT being computed, the same parameters amsubscript𝑎𝑚a_{m}italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT are employed, except that the first iteration with a1=0subscript𝑎10a_{1}=0italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0 (which gives the explicit Euler solution) is not carried out. One can show [58] that formulas (3.17) can be written in an operator form as

𝐜n+1=(𝐈𝐅p2)[𝐈+τ𝐀^]1𝐟^n,superscript𝐜𝑛1𝐈superscriptsubscript𝐅𝑝2superscriptdelimited-[]𝐈τ^𝐀1superscript^𝐟𝑛\mathbf{c}^{n+1}=(\mathbf{I}-\mathbf{F}_{p}^{2})\left[\mathbf{I}+\tauup% \widehat{\mathbf{A}}\right]^{-1}\widehat{\mathbf{f}}^{n},bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT = ( bold_I - bold_F start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) [ bold_I + roman_τ over^ start_ARG bold_A end_ARG ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT over^ start_ARG bold_f end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT , (3.18)

where 𝐀^^𝐀\widehat{\mathbf{A}}over^ start_ARG bold_A end_ARG and 𝐟^nsuperscript^𝐟𝑛\widehat{\mathbf{f}}^{n}over^ start_ARG bold_f end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT are defined in (3.15) and 𝐅psubscript𝐅𝑝\mathbf{F}_{p}bold_F start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the Chebyshev polynomial operator,

𝐅p=m=pm=1(𝐈11+τam(𝐈+τ𝐀^)).subscript𝐅𝑝superscriptsubscriptproduct𝑚𝑝𝑚1𝐈11τsubscript𝑎𝑚𝐈τ^𝐀\mathbf{F}_{p}=\prod_{m=p}^{m=1}\left(\mathbf{I}-\frac{1}{1+\tauup a_{m}}(% \mathbf{I}+\tauup\widehat{\mathbf{A}})\right).bold_F start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = ∏ start_POSTSUBSCRIPT italic_m = italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m = 1 end_POSTSUPERSCRIPT ( bold_I - divide start_ARG 1 end_ARG start_ARG 1 + roman_τ italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG ( bold_I + roman_τ over^ start_ARG bold_A end_ARG ) ) .

By replacing 𝐅p2superscriptsubscript𝐅𝑝2\mathbf{F}_{p}^{2}bold_F start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT in (3.18) by 𝐅psubscript𝐅𝑝\mathbf{F}_{p}bold_F start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT we obtain an operator form of the (regular) LI scheme. Thus, as we see, squaring the Chebyshev polynomial leads to monotonicity in the LIM scheme.

Since scheme (3.17) under consideration is derived from the linearly stabilized splitting (3.14), we call this scheme LIM-LSS. To discover the Eyre splitting effect in this scheme, we include another scheme the numerical experiments presented below. It is obtained by linearizing implicit Euler scheme (3.10). We linearize the scheme by approximating the nonlinear implicit scheme in (3.10) as

F(𝐜n+1)F(𝐜n)+Jn(𝐜n+1𝐜n),superscript𝐹superscript𝐜𝑛1superscript𝐹superscript𝐜𝑛subscript𝐽𝑛superscript𝐜𝑛1superscript𝐜𝑛F^{\prime}(\mathbf{c}^{n+1})\approx F^{\prime}(\mathbf{c}^{n})+J_{n}(\mathbf{c% }^{n+1}-\mathbf{c}^{n}),italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ≈ italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) + italic_J start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT - bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) , (3.19)

where Jnsubscript𝐽𝑛J_{n}italic_J start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the Jacobian of the mapping Fsuperscript𝐹F^{\prime}italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT evaluated at 𝐜nsuperscript𝐜𝑛\mathbf{c}^{n}bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT. Substituting approximation (3.19) in (3.10) leads to a linearized implicit Euler (LIE) scheme

𝐜n+1=[𝐈+τ𝐀~n]1𝐟~n,𝐀~=𝐀(Jn+ϵ2𝐀),𝐟~n=𝐜n+τ𝐀(Jn𝐜nF(𝐜n)),formulae-sequencesuperscript𝐜𝑛1superscriptdelimited-[]𝐈τsubscript~𝐀𝑛1superscript~𝐟𝑛formulae-sequence~𝐀𝐀subscript𝐽𝑛superscriptϵ2𝐀superscript~𝐟𝑛superscript𝐜𝑛τ𝐀subscript𝐽𝑛superscript𝐜𝑛superscript𝐹superscript𝐜𝑛\mathbf{c}^{n+1}=\left[\mathbf{I}+\tauup\widetilde{\mathbf{A}}_{n}\right]^{-1}% \widetilde{\mathbf{f}}^{n},\quad\widetilde{\mathbf{A}}=\mathbf{A}(J_{n}+% \epsilonup^{2}\mathbf{A}),\quad\widetilde{\mathbf{f}}^{n}=\mathbf{c}^{n}+% \tauup\mathbf{A}(J_{n}\mathbf{c}^{n}-F^{\prime}(\mathbf{c}^{n})),bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT = [ bold_I + roman_τ over~ start_ARG bold_A end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT over~ start_ARG bold_f end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT , over~ start_ARG bold_A end_ARG = bold_A ( italic_J start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT + roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_A ) , over~ start_ARG bold_f end_ARG start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT + roman_τ bold_A ( italic_J start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ) , (3.20)

where, just as in scheme (3.15), a linear system solution is understood, so that the inverse matrix is not computed. Comparing schemes (3.15) and (3.20), one can notice that replacing Jnsubscript𝐽𝑛J_{n}italic_J start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT in (3.20) by 2𝐈2𝐈2\mathbf{I}2 bold_I leads to scheme (3.15).

The LIM scheme based on (3.20) can be obtained in exactly the same as the LIM scheme based on (3.15). One can simply repeat the derivations given above, replacing 𝐀^^𝐀\widehat{\mathbf{A}}over^ start_ARG bold_A end_ARG and 𝐟^nsubscript^𝐟𝑛\widehat{\mathbf{f}}_{n}over^ start_ARG bold_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT by 𝐀~nsubscript~𝐀𝑛\widetilde{\mathbf{A}}_{n}over~ start_ARG bold_A end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT and 𝐟~nsubscript~𝐟𝑛\widetilde{\mathbf{f}}_{n}over~ start_ARG bold_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, respectively. We call the LIM scheme derived in this way the LIM-LIE (LIM linearized implicit Euler) scheme.

A computational efficiency of the LIM schemes can be established, based on the number of required iterations (3.16) and the following reasoning [54, 58]. If the time interval for which the problem has to be solved is increased by a factor s𝑠sitalic_s then computational costs of an explicit scheme should also grow by a factor s𝑠sitalic_s (as the number of required time steps is s𝑠sitalic_s times larger). The same increase in computational costs is also observed if the upper spectral bound λsubscriptλ\lambdaup_{\infty}roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT is increased by a factor s𝑠sitalic_s (since a time step size in an explicit scheme is usually bound to a stability condition of the form τ2/λτ2subscriptλ\tauup\leqslant 2/\lambdaup_{\infty}roman_τ ⩽ 2 / roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT). Recall that for the Cahn-Hilliard equation, in general, we have λh4similar-tosubscriptλsuperscript4\lambdaup_{\infty}\sim h^{-4}roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ∼ italic_h start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT. In the LIM schemes the costs grow slower: an increase of the time interval or the upper spectral bound λsubscriptλ\lambdaup_{\infty}roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT by a factor s𝑠sitalic_s leads to an increase in computational work approximately by a factor s𝑠\sqrt{s}square-root start_ARG italic_s end_ARG (as the number of required Chebyshev iterations p(τλ)1/2similar-to𝑝superscriptτsubscriptλ12p\sim(\tauup\lambdaup_{\infty})^{1/2}italic_p ∼ ( roman_τ roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT, see (3.16)). These estimates make clear that the gain in computational costs provided by the LIM schemes increases with the problem size N=1/h𝑁1N=1/hitalic_N = 1 / italic_h [54, 58]. Nevertheless, it should also be taken into account that the gain actually observed in practice can be restricted by other factors, such as accuracy requirements.

4.  Computational experiments

4.1.  Gradient stability tests

In the tests presented here initial-value problem (2.1)–(2.3) is solved for the 1D Cahn-Hilliard equation. Space discretization is done by the standard second order finite differences on a uniform grid, as discussed in the beginning of Section 3.1. This reduces the original initial-boundary value problem to an initial value problem (3.8) to be solved by the time integration schemes (3.9)–(3.13) and new local iteration schemes (3.17) based on (3.14) and (3.20) (the LIM-LSS and LIM-LIE schemes). The parameter ϵϵ\epsilonuproman_ϵ is set to one of the following two values:

ϵϵ\displaystyle\epsilonuproman_ϵ =ϵ4(h),ϵm(h)hm22arth(9/10),formulae-sequenceabsentsubscriptϵ4subscriptϵ𝑚𝑚22arth910\displaystyle=\epsilonup_{4}(h),\quad\epsilonup_{m}(h)\equiv\frac{hm}{2\sqrt{2% }\,\mathrm{arth}(9/10)},= roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h ) , roman_ϵ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( italic_h ) ≡ divide start_ARG italic_h italic_m end_ARG start_ARG 2 square-root start_ARG 2 end_ARG roman_arth ( 9 / 10 ) end_ARG , (4.1)
ϵϵ\displaystyle\epsilonuproman_ϵ =ϵ4(1/64),absentsubscriptϵ4164\displaystyle=\epsilonup_{4}(1/64),= roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( 1 / 64 ) , (4.2)

i.e., ϵϵ\epsilonuproman_ϵ either depends on the space grid size hhitalic_h (formula (4.1), or set to a fixed value ϵ=ϵ4(1/64)ϵsubscriptϵ4164\epsilonup=\epsilonup_{4}(1/64)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( 1 / 64 ) of the grid N=64𝑁64N=64italic_N = 64 (formula (4.2)). Note that m=4𝑚4m=4italic_m = 4 in formula (4.1) defines a characteristic number of the grid cells on which a stationary solution changes from its minimum to maximum values.

For all the tested time integration schemes, largest time step sizes for which the gradient stability is observed are reported in Table 2. In these runs, the grid-dependent ϵϵ\epsilonuproman_ϵ values (4.1) are used and a scheme is considered to be gradient stable provided that the discrete energy (3.7) does not increase more than 1% each time step, i.e., for each time step n𝑛nitalic_n holds

ψh(𝐜n+1)1.01ψh(𝐜n).subscriptψsuperscript𝐜𝑛11.01subscriptψsuperscript𝐜𝑛\psiup_{h}(\mathbf{c}^{n+1})\leqslant 1.01\psiup_{h}(\mathbf{c}^{n}).roman_ψ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ⩽ 1.01 roman_ψ start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) . (4.3)

Gradient stability is tested on several initial value vectors 𝐜0superscript𝐜0\mathbf{c}^{0}bold_c start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT, where each vector component is taken to be an independent and identically distributed random variable in [0,1]01[0,1][ 0 , 1 ], rounded up to the second digit after the decimal point. Table 2 also contains results for the linearized implicit Euler (LIE) scheme (3.20).

As can be seen in the table, only Eyre splitting based schemes turn out to be unconditionally gradient stable. In the tests, a scheme is considered to be unconditionally gradient stable if gradient stability is observed for τ500τ500\tauup\leqslant 500roman_τ ⩽ 500. Furthermore, we note that maximum time step size in the implicit scheme exceeds the maximum time step size in the explicit schemes approximately by the same factor (60absent60\approx 60≈ 60) for all hhitalic_h. Comparing the LIE and IE schemes, we see that linearization does not corrupt the gradient stability of the implicit scheme. The LIM-LSS scheme conserves the unconditional gradient stability of the LSS scheme, and the LIM-LIE scheme inherits the conditional gradient stability of the LIE scheme. One of the aims of this work is to show that implicit schemes can be successfully replaced by stable explicit schemes. Having this in mind, we may conclude from Table 2 that both LIM schemes work well. This, of course, does imply that increasing the time step size ττ\tauuproman_τ in both LIM-LSS and LIM-LIE schemes leads to an increase of the Chebyshev iterations required per time step. We also note that the values presented in our Table 2 are close to the values reported in Table 1 from paper [64].

Table 2: Largest time step size ττ\tauuproman_τ for which time integration schemes are gradient stable, i.e., condition (4.3) holds for ϵ=ϵ4(h)ϵsubscriptϵ4\epsilonup=\epsilonup_{4}(h)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h )
hhitalic_h 1/321321/321 / 32 1/641641/641 / 64 1/12811281/1281 / 128 1/25612561/2561 / 256
EE 8.8×1058.8superscript1058.8\times 10^{-5}8.8 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT 2.1×1052.1superscript1052.1\times 10^{-5}2.1 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT 5.4×1065.4superscript1065.4\times 10^{-6}5.4 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT 1.3×1061.3superscript1061.3\times 10^{-6}1.3 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT
IE 5.1×1035.1superscript1035.1\times 10^{-3}5.1 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 1.3×1031.3superscript1031.3\times 10^{-3}1.3 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 3.0×1043.0superscript1043.0\times 10^{-4}3.0 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT 8.1×1058.1superscript1058.1\times 10^{-5}8.1 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
CN 2.4×1032.4superscript1032.4\times 10^{-3}2.4 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 6.8×1046.8superscript1046.8\times 10^{-4}6.8 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT 1.7×1041.7superscript1041.7\times 10^{-4}1.7 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT 3.8×1053.8superscript1053.8\times 10^{-5}3.8 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
SIE 2.2×1032.2superscript1032.2\times 10^{-3}2.2 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 5.7×1045.7superscript1045.7\times 10^{-4}5.7 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT 1.9×1041.9superscript1041.9\times 10^{-4}1.9 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT 4.0×1054.0superscript1054.0\times 10^{-5}4.0 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
LSS \infty \infty \infty \infty
NLSS \infty \infty \infty \infty
LIM-LSS \infty \infty \infty \infty
LIM-LIE 9.9×1039.9superscript1039.9\times 10^{-3}9.9 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 2.8×1032.8superscript1032.8\times 10^{-3}2.8 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 5.8×1045.8superscript1045.8\times 10^{-4}5.8 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT 1.2×1041.2superscript1041.2\times 10^{-4}1.2 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
LIE 5.8×1035.8superscript1035.8\times 10^{-3}5.8 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 1.6×1031.6superscript1031.6\times 10^{-3}1.6 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT 3.2×1043.2superscript1043.2\times 10^{-4}3.2 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT 8.3×1058.3superscript1058.3\times 10^{-5}8.3 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT

4.2.  Accuracy and efficiency tests

The tests discussed here are aimed to evaluate accuracy and computational efficiency of the considered schemes and, in particular, to discover whether the LIM schemes provide a gain in computational costs with respect to the explicit Euler (EE) scheme. Since the costs per time step in the EE scheme are minimal, it is clear that other time integration schemes can deliver a higher computational efficiency only if the time step size is increased. However, increasing the time step size is possible only provided the delivered numerical accuracy remains within the allowed tolerance. We estimate accuracy of a time integration scheme by comparing its solution at the final time t=T𝑡𝑇t=Titalic_t = italic_T to the reference solution 𝐜ref(T)subscript𝐜ref𝑇\mathbf{c}_{\text{ref}}(T)bold_c start_POSTSUBSCRIPT ref end_POSTSUBSCRIPT ( italic_T ), which is computed for each space grid with a tiny time step size τ=109τsuperscript109\tauup=10^{-9}roman_τ = 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT. The initial value vector is set to have random entries, in the same way as in the previous tests, then fixed and used for a current space grid in all time integration schemes. For each scheme the achieved accuracy is measured as a relative error norm

𝐜n𝐜ref(T)𝐜ref(T),n=nfinal=T/τ,normsuperscript𝐜𝑛subscript𝐜ref𝑇normsubscript𝐜ref𝑇𝑛subscript𝑛final𝑇τ\frac{\|\mathbf{c}^{n}-\mathbf{c}_{\text{ref}}(T)\|}{\|\mathbf{c}_{\text{ref}}% (T)\|},\quad n=n_{\text{final}}=T/\tauup,divide start_ARG ∥ bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - bold_c start_POSTSUBSCRIPT ref end_POSTSUBSCRIPT ( italic_T ) ∥ end_ARG start_ARG ∥ bold_c start_POSTSUBSCRIPT ref end_POSTSUBSCRIPT ( italic_T ) ∥ end_ARG , italic_n = italic_n start_POSTSUBSCRIPT final end_POSTSUBSCRIPT = italic_T / roman_τ , (4.4)

where 𝐜=𝐜T𝐜norm𝐜superscript𝐜𝑇𝐜\|\mathbf{c}\|=\sqrt{\mathbf{c}^{T}\mathbf{c}}∥ bold_c ∥ = square-root start_ARG bold_c start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_c end_ARG is the Euclidean vector norm.

In all tests we set the final time T𝑇Titalic_T equal to T=0.2𝑇0.2T=0.2italic_T = 0.2. At this T𝑇Titalic_T the solution has already passed the initial phase of forming homogeneous regions for times tϵ2𝑡superscriptϵ2t\approx\epsilonup^{2}italic_t ≈ roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, all the schemes need sufficiently many time steps, but the solution is still far from its stationary state.

Table 3 presents the values of the upper spectral bound λsubscriptλ\lambdaup_{\infty}roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT depending on the grid step hhitalic_h and the corresponding maximum time step size values ττ\tauuproman_τ at which the explicit scheme is stable. In each case, two maximum ττ\tauuproman_τ values are given in the table:
(a) a maximum time step size ττ\tauuproman_τ, for which gradient stability is observed (i.e., no energy increase (4.3) takes place); (b) a maximum time step size ττ\tauuproman_τ, for which the schemes LIM-LIE and LIM-LSS operate in the explicit scheme mode, i.e., for which relation (3.16) yields p=1𝑝1p=1italic_p = 1, meaning that no Chebyshev iterations are needed.
The last condition implies that the EE scheme is stable in the Euclidean operator norm. This can be seen from the following. If for the LIM-LIE scheme relation (3.16) gives p=1𝑝1p=1italic_p = 1 then

τλ(4π)21,τsubscriptλsuperscript4π21\tauup\lambdaup_{\infty}\leqslant\left(\frac{4}{\piup}\right)^{2}-1,roman_τ roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT ⩽ ( divide start_ARG 4 end_ARG start_ARG roman_π end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 , (4.5)

where λ=𝐀~1subscriptλsubscriptnorm~𝐀1\lambdaup_{\infty}=\|\widetilde{\mathbf{A}}\|_{1}roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = ∥ over~ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT with 𝐀~~𝐀\widetilde{\mathbf{A}}over~ start_ARG bold_A end_ARG being defined in (3.20). If Jn=F(𝐜n)/𝐜subscript𝐽𝑛superscript𝐹superscript𝐜𝑛𝐜J_{n}=\partial F^{\prime}(\mathbf{c}^{n})/\partial\mathbf{c}italic_J start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = ∂ italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) / ∂ bold_c then, taking into account approximation F(𝐜n)F(0)+Jn(𝐜n0)=Jn𝐜nsuperscript𝐹superscript𝐜𝑛superscript𝐹0subscript𝐽𝑛superscript𝐜𝑛0subscript𝐽𝑛subscript𝐜𝑛F^{\prime}(\mathbf{c}^{n})\approx F^{\prime}(\textbf{0})+J_{n}(\mathbf{c}^{n}-% \textbf{0})=J_{n}\mathbf{c}_{n}italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ≈ italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 0 ) + italic_J start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - 0 ) = italic_J start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT bold_c start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, we obtain, for the EE solution 𝐜n+1superscript𝐜𝑛1\mathbf{c}^{n+1}bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT,

𝐜n+1=𝐜nτ𝐀(F(𝐜n)+ϵ2𝐀𝐜n)(Iτ𝐀~)𝐜nIτ𝐀~𝐜n.normsuperscript𝐜𝑛1normsuperscript𝐜𝑛τ𝐀superscript𝐹superscript𝐜𝑛superscriptϵ2superscript𝐀𝐜𝑛norm𝐼τ~𝐀superscript𝐜𝑛norm𝐼τ~𝐀normsuperscript𝐜𝑛\|\mathbf{c}^{n+1}\|=\|\mathbf{c}^{n}-\tauup\mathbf{A}(F^{\prime}(\mathbf{c}^{% n})+\epsilonup^{2}\mathbf{A}\mathbf{c}^{n})\|\approx\|(I-\tauup\widetilde{% \mathbf{A}})\mathbf{c}^{n}\|\leqslant\|I-\tauup\widetilde{\mathbf{A}}\|\,\|% \mathbf{c}^{n}\|.∥ bold_c start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ∥ = ∥ bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT - roman_τ bold_A ( italic_F start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) + roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_Ac start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ∥ ≈ ∥ ( italic_I - roman_τ over~ start_ARG bold_A end_ARG ) bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ∥ ⩽ ∥ italic_I - roman_τ over~ start_ARG bold_A end_ARG ∥ ∥ bold_c start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ∥ .

Since 𝐀~~𝐀\widetilde{\mathbf{A}}over~ start_ARG bold_A end_ARG is symmetric positive semidefinite matrix, condition Iτ𝐀~21subscriptnorm𝐼τ~𝐀21\|I-\tauup\widetilde{\mathbf{A}}\|_{2}\leqslant 1∥ italic_I - roman_τ over~ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⩽ 1 in the Euclidean operator norm is equivalent to τ𝐀~22τsubscriptnorm~𝐀22\tauup\|\widetilde{\mathbf{A}}\|_{2}\leqslant 2roman_τ ∥ over~ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⩽ 2, which follows from (4.5) (note that 𝐀~2𝐀~1=λsubscriptnorm~𝐀2subscriptnorm~𝐀1subscriptλ\|\widetilde{\mathbf{A}}\|_{2}\leqslant\|\widetilde{\mathbf{A}}\|_{1}=% \lambdaup_{\infty}∥ over~ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⩽ ∥ over~ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT).

Table 3: The upper spectral bound λ=𝐀^1subscriptλsubscriptnorm^𝐀1\lambdaup_{\infty}=\|\widehat{\mathbf{A}}\|_{1}roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = ∥ over^ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT for the LIM-LSS scheme and the maximum time step size τmaxsubscriptτ\tauup_{\max}roman_τ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT, for which the EE scheme is gradient stable (row 3 in the table) and the LIM-LSS scheme operates in the explicit scheme mode (rows 4 and 6), depending on hhitalic_h. As seen in the table, λ=𝒪(h2)subscriptλ𝒪superscript2\lambdaup_{\infty}=\mathcal{O}(h^{-2})roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = caligraphic_O ( italic_h start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ) for ϵ=ϵ4(h)ϵsubscriptϵ4\epsilonup=\epsilonup_{4}(h)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h ), λ=𝒪(h4)subscriptλ𝒪superscript4\lambdaup_{\infty}=\mathcal{O}(h^{-4})roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = caligraphic_O ( italic_h start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT ) for ϵ=ϵ4(1/64)ϵsubscriptϵ4164\epsilonup=\epsilonup_{4}(1/64)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( 1 / 64 ) and in all cases τmax=𝒪(λ1)subscriptτ𝒪superscriptsubscriptλ1\tauup_{\max}=\mathcal{O}(\lambdaup_{\infty}^{-1})roman_τ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = caligraphic_O ( roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ).
hhitalic_h 1/321321/321 / 32 1/641641/641 / 64 1/12811281/1281 / 128 1/25612561/2561 / 256 1/51215121/5121 / 512
ϵ=ϵ4(h)ϵsubscriptϵ4\epsilonup=\epsilonup_{4}(h)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h )
λ=𝐀^1subscriptλsubscriptnorm^𝐀1\lambdaup_{\infty}=\|\widehat{\mathbf{A}}\|_{1}roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = ∥ over^ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT 2.3×1042.3superscript1042.3\times 10^{4}2.3 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 9.3×1049.3superscript1049.3\times 10^{4}9.3 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 3.7×1053.7superscript1053.7\times 10^{5}3.7 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT 1.5×1061.5superscript1061.5\times 10^{6}1.5 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT 6.0×1066.0superscript1066.0\times 10^{6}6.0 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT
gradient stability 8.8×1058.8superscript1058.8\times 10^{-5}8.8 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT 2.1×1052.1superscript1052.1\times 10^{-5}2.1 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT 5.4×1065.4superscript1065.4\times 10^{-6}5.4 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT 1.3×1061.3superscript1061.3\times 10^{-6}1.3 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT 3.3×1073.3superscript1073.3\times 10^{-7}3.3 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT
explicit scheme mode 2.6×1052.6superscript1052.6\times 10^{-5}2.6 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT 6.6×1066.6superscript1066.6\times 10^{-6}6.6 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT 1.6×1061.6superscript1061.6\times 10^{-6}1.6 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT 4.1×1074.1superscript1074.1\times 10^{-7}4.1 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT 1.0×1071.0superscript1071.0\times 10^{-7}1.0 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT
ϵ=ϵ4(1/64)ϵsubscriptϵ4164\epsilonup=\epsilonup_{4}(1/64)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( 1 / 64 )
λ=𝐀^1subscriptλsubscriptnorm^𝐀1\lambdaup_{\infty}=\|\widehat{\mathbf{A}}\|_{1}roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = ∥ over^ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT 1.2×1041.2superscript1041.2\times 10^{4}1.2 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 9.3×1049.3superscript1049.3\times 10^{4}9.3 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT 1.1×1061.1superscript1061.1\times 10^{6}1.1 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT 1.6×1071.6superscript1071.6\times 10^{7}1.6 × 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT 2.5×1082.5superscript1082.5\times 10^{8}2.5 × 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT
explicit scheme mode 5.1×1055.1superscript1055.1\times 10^{-5}5.1 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT 6.6×1066.6superscript1066.6\times 10^{-6}6.6 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT 5.6×1075.6superscript1075.6\times 10^{-7}5.6 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT 3.8×1083.8superscript1083.8\times 10^{-8}3.8 × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT 2.4×1092.4superscript1092.4\times 10^{-9}2.4 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT

Table 3 shows dependence of the upper spectral bound λsubscriptλ\lambdaup_{\infty}roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT on hhitalic_h for the LIM-LSS scheme, i.e., λ=𝐀^1subscriptλsubscriptnorm^𝐀1\lambdaup_{\infty}=\|\widehat{\mathbf{A}}\|_{1}roman_λ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = ∥ over^ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. In the tests presented here in all cases the values of 𝐀^1subscriptnorm^𝐀1\|\widehat{\mathbf{A}}\|_{1}∥ over^ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and 𝐀~1subscriptnorm~𝐀1\|\widetilde{\mathbf{A}}\|_{1}∥ over~ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT appear to be so close that the Chebyshev iteration numbers in the LIM-LSS and LIM-LIE schemes turn out to be almost the same. The values in Table 3 are given for the both ways to determine ϵϵ\epsilonuproman_ϵ, see (4.1), (4.2). From the table data, it is easy to check that for the choice ϵ=ϵ4(h)ϵsubscriptϵ4\epsilonup=\epsilonup_{4}(h)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h ) the matrix ϵ2𝐀superscriptϵ2𝐀\epsilonup^{2}\mathbf{A}roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_A does not depend on hhitalic_h. Hence, taking into account (3.15), we obtain a dependence 𝐀^1=𝒪(h2)subscriptnorm^𝐀1𝒪superscript2\|\widehat{\mathbf{A}}\|_{1}=\mathcal{O}(h^{-2})∥ over^ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = caligraphic_O ( italic_h start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ). For ϵϵ\epsilonuproman_ϵ defined in the second way (by formula (4.2)), we have 𝐀^1=𝒪(h4)subscriptnorm^𝐀1𝒪superscript4\|\widehat{\mathbf{A}}\|_{1}=\mathcal{O}(h^{-4})∥ over^ start_ARG bold_A end_ARG ∥ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = caligraphic_O ( italic_h start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT ).

Refer to caption
Figure 2: Convergence (achieved accuracy versus the time step size ττ\tauuproman_τ) for the LIM-LSS, LSS, LIM-LIE and LIE schemes on the N=256𝑁256N=256italic_N = 256 space grid, ϵ=ϵ4(h)ϵsubscriptϵ4\epsilonup=\epsilonup_{4}(h)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h )

The first order convergence 𝒪(τ)𝒪τ\mathcal{O}(\tauup)caligraphic_O ( roman_τ ) of the schemes LIM-LSS, LSS, LIM-LIE and LIE is confirmed by the plot in Figure 2. As can be seen from the figure, both LIM schemes retain the accuracy properties of the of the schemes they are based on. In addition, we note that the Eyre splitting schemes LSS and LIM-LSS turn out to be less accurate than the LIM-LIE and LIE schemes. In the former, the gradient stability is achieved by the splitting, which leads to an additional error.

In Tables 47 computational costs and achieved accuracies, depending on the time step size ττ\tauuproman_τ, are given for the EE, LSS, LIM-LSS, and LIM-LIE schemes. As we see in tables, accuracies delivered by the LSS and LIM-LSS schemes turn out to be too low (we consider error values (4.4) greater than 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT unacceptably large). In these schemes, a time step size increase with respect to the EE scheme, needed to compensate higher costs per time step, leads to an unacceptably low accuracy. The LIM-LIE scheme is more accurate and yields a gain with respect to the EE scheme on fine grids (the gain is approximately a factor 2 on the N=512𝑁512N=512italic_N = 512 grid). The fact that the low accuracy of the LSS and LIM-LSS is caused by the Eyre splitting, is confirmed by switching to the LIE and LIM-LIE schemes (the LIE scheme error values are not shown in the table, they are close to that of the LIM-LIE scheme, see Figure 2).

Table 4: Number of matrix-vector multiplications (matvecs), number of linear system solutions (for implicit schemes LSS and LIE), and achieved accuracy versus ττ\tauuproman_τ. ϵ=ϵ4(h)ϵsubscriptϵ4\epsilonup=\epsilonup_{4}(h)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h ), space grid N=64𝑁64N=64italic_N = 64
ττ\tauuproman_τ scheme # matvecs /// error
# lin.syst. solutions ×103absentsuperscript103\times 10^{3}× 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (4.4)
5.0×1055.0superscript1055.0\times 10^{-5}5.0 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT LIM-LSS 12 / — 5.24×1025.24superscript1025.24\times 10^{-2}5.24 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LSS 4 / 4 1.18×1011.18superscript1011.18\times 10^{-1}1.18 × 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT
LIM-LIE 12 / — 8.34×1038.34superscript1038.34\times 10^{-3}8.34 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT
1.0×1051.0superscript1051.0\times 10^{-5}1.0 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT EE 20 / — 1.04×1041.04superscript1041.04\times 10^{-4}1.04 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
LIM-LSS 60 / — 2.52×1022.52superscript1022.52\times 10^{-2}2.52 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LSS 20 / 20 2.92×1022.92superscript1022.92\times 10^{-2}2.92 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LIM-LIE 60 / — 3.96×1043.96superscript1043.96\times 10^{-4}3.96 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
1.0×1061.0superscript1061.0\times 10^{-6}1.0 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT EE 200 / — 2.27×1052.27superscript1052.27\times 10^{-5}2.27 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
LIM-LSS 200 / — 2.27×1052.27superscript1052.27\times 10^{-5}2.27 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
LSS 200 / 200 2.93×1032.93superscript1032.93\times 10^{-3}2.93 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT
LIM-LIE 200 / — 2.27×1052.27superscript1052.27\times 10^{-5}2.27 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
Table 5: Number of matrix-vector multiplications (matvecs), number of linear system solutions (for implicit schemes LSS and LIE), and achieved accuracy versus ττ\tauuproman_τ. ϵ=ϵ4(h)ϵsubscriptϵ4\epsilonup=\epsilonup_{4}(h)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h ), space grid N=128𝑁128N=128italic_N = 128.
ττ\tauuproman_τ scheme # matvecs /// error
# lin.syst. solutions ×103absentsuperscript103\times 10^{3}× 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (4.4)
2.5×1052.5superscript1052.5\times 10^{-5}2.5 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT LIM-LSS 40 / — 1.71×1011.71superscript1011.71\times 10^{-1}1.71 × 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT
LSS 8 / 8 1.66×1011.66superscript1011.66\times 10^{-1}1.66 × 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT
LIM-LIE 40 / — 9.06×1029.06superscript1029.06\times 10^{-2}9.06 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
5.0×1065.0superscript1065.0\times 10^{-6}5.0 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT EE 40 / — 3.37×1043.37superscript1043.37\times 10^{-4}3.37 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
LIM-LSS 120 / — 3.60×1023.60superscript1023.60\times 10^{-2}3.60 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LSS 40 / 40 1.19×1021.19superscript1021.19\times 10^{-2}1.19 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LIM-LIE 120 / — 7.91×1047.91superscript1047.91\times 10^{-4}7.91 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
Table 6: Number of matrix-vector multiplications (matvecs), number of linear system solutions (for implicit schemes LSS and LIE), and achieved accuracy versus ττ\tauuproman_τ. ϵ=ϵ4(h)ϵsubscriptϵ4\epsilonup=\epsilonup_{4}(h)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h ), space grid N=256𝑁256N=256italic_N = 256.
ττ\tauuproman_τ scheme # matvecs /// error
# lin.syst. solutions ×103absentsuperscript103\times 10^{3}× 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (4.4)
1.0×1051.0superscript1051.0\times 10^{-5}1.0 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT LIM-LIE 140 / — 8.14×1028.14superscript1028.14\times 10^{-2}8.14 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
5.0×1065.0superscript1065.0\times 10^{-6}5.0 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT LIM-LSS 200 / — 4.49×1024.49superscript1024.49\times 10^{-2}4.49 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LSS 40 / 40 8.12×1028.12superscript1028.12\times 10^{-2}8.12 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LIM-LIE 200 / — 1.44×1041.44superscript1041.44\times 10^{-4}1.44 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
1.0×1061.0superscript1061.0\times 10^{-6}1.0 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT EE 200 / — 2.06×1052.06superscript1052.06\times 10^{-5}2.06 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
LIM-LSS 600 / — 4.48×1024.48superscript1024.48\times 10^{-2}4.48 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LSS 200 / 200 4.48×1024.48superscript1024.48\times 10^{-2}4.48 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LIM-LIE 600 / — 1.72×1051.72superscript1051.72\times 10^{-5}1.72 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
Table 7: Number of matrix-vector multiplications (matvecs), number of linear system solutions (for implicit schemes LSS and LIE), and achieved accuracy versus ττ\tauuproman_τ, ϵ=ϵ4(h)ϵsubscriptϵ4\epsilonup=\epsilonup_{4}(h)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h ), space grid N=512𝑁512N=512italic_N = 512.
ττ\tauuproman_τ scheme # matvecs /// error
# lin.syst. solutions ×106absentsuperscript106\times 10^{6}× 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT (4.4)
2.0×1062.0superscript1062.0\times 10^{-6}2.0 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT LIM-LIE 0.5 / — 8.25×1068.25superscript1068.25\times 10^{-6}8.25 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT
1.0×1061.0superscript1061.0\times 10^{-6}1.0 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT LIM-LSS 1 / — 3.25×1023.25superscript1023.25\times 10^{-2}3.25 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LSS 0.2 / 0.2 3.25×1023.25superscript1023.25\times 10^{-2}3.25 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LIM-LIE 1 / — 4.61×1064.61superscript1064.61\times 10^{-6}4.61 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT
2.0×1072.0superscript1072.0\times 10^{-7}2.0 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT EE 1 / — 9.57×1079.57superscript1079.57\times 10^{-7}9.57 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT
LIM-LSS 3 / — 3.25×1023.25superscript1023.25\times 10^{-2}3.25 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LSS 1 / 1 3.25×1023.25superscript1023.25\times 10^{-2}3.25 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LIM-LIE 3 / — 7.35×1077.35superscript1077.35\times 10^{-7}7.35 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT

Plots showing achieved accuracy versus computational costs are given in Figure 3. As can be seen in the plots, the LIM schemes perform better on the N=512𝑁512N=512italic_N = 512 grid than on the N=256𝑁256N=256italic_N = 256 grid: a cost reduction of an approximately factor 3 is achieved on the N=512𝑁512N=512italic_N = 512 grid.

Refer to caption
Refer to caption
Figure 3: Achieved accuracy versus number of matrix-vector multiplications (matvecs) for the EE and LIM schemes on the N=256𝑁256N=256italic_N = 256 space grid (upper plot) and N=512𝑁512N=512italic_N = 512 space grid (lower plot), ϵ=ϵ4(h)ϵsubscriptϵ4\epsilonup=\epsilonup_{4}(h)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h ). Increasing ττ\tauuproman_τ in the EE scheme further is impossible due to the stability restrictions.

In the presented tests a rather low achieved accuracy is observed for all the schemes except the EE scheme and this is not only due to the Eyre splitting. In our tests the initial value vector is a non-smooth grid function, where each vector entry is chosen randomly. As discussed above in the introduction, a fast forming of the homogeneity regions is observed for such initial values, taking place at typical evolution times ϵ2superscriptϵ2\epsilonup^{2}roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Hence, to properly track this process, the time step size should be chosen as τ𝒪(ϵ2)=𝒪(h2)similar-toτ𝒪superscriptϵ2𝒪superscript2\tauup\sim\mathcal{O}(\epsilonup^{2})=\mathcal{O}(h^{2})roman_τ ∼ caligraphic_O ( roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) = caligraphic_O ( italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). Therefore, taking into account that ϵ2Aϵ2h2=𝒪(1)similar-tosuperscriptϵ2norm𝐴superscriptϵ2superscript2𝒪1\epsilonup^{2}\|A\|\sim\epsilonup^{2}h^{-2}=\mathcal{O}(1)roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∥ italic_A ∥ ∼ roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_h start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT = caligraphic_O ( 1 ), we see that during this initial phase of the time integration the time step size ττ\tauuproman_τ can not be taken much larger than the time step size in the EE scheme. This time step is, in this case, determined by the accuracy rather than by the stability requirements.

To test the stability and accuracy of the considered time integration schemes at larger time steps, we carry out tests with a smoother initial condition. For the space grids N>64𝑁64N>64italic_N > 64, we set it by interpolating the initial non-smooth initial solution from the grid N=64𝑁64N=64italic_N = 64 to the current (finer) space grid. On the N=64𝑁64N=64italic_N = 64 space grid the initial vector does not change. The interpolation is done by the piecewise cubic Hermite interpolation, which excludes the occurrence of new extrema and yields a continuously differentiable function (in the octave package, this interpolation method is called pchip). In addition, in the tests with the smoothed initial value vector, the second method (4.2) for setting ϵϵ\epsilonuproman_ϵ is employed, which allows to test our schemes in the situation when the norm of the right-hand side operator grows as 𝒪(h4)𝒪superscript4\mathcal{O}(h^{-4})caligraphic_O ( italic_h start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT ), see Table 3.

Refer to caption
Figure 4: Convergence (achieved accuracy versus the time step size ττ\tauuproman_τ) for the LIM-LSS, LSS, LIM-LIE and LIE schemes on the N=256𝑁256N=256italic_N = 256 space grid, ϵ=ϵ4(1/64)ϵsubscriptϵ4164\epsilonup=\epsilonup_{4}(1/64)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( 1 / 64 )

We start by testing the first order convergence 𝒪(τ)𝒪τ\mathcal{O}(\tauup)caligraphic_O ( roman_τ ) at the plot in Figure 4. Comparing the plot with the plot in 2, we see that all the schemes achieve a much higher accuracy. However, the LSS and LIM-LSS schemes based on the Eyre splitting are still less accurate than the LIE and LIM-LIE schemes. Small error oscillations observed with the LIM-LIE scheme for τ106τsuperscript106\tauup\approx 10^{-6}roman_τ ≈ 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT are caused by the fact that number of Chebyshev iterations carried out each time step slightly varies for these ττ\tauuproman_τ values.

Test results for smoothed initial value vectors and fixed ϵ=ϵ4(1/64)ϵsubscriptϵ4164\epsilonup=\epsilonup_{4}(1/64)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( 1 / 64 ) are presented in Tables 810 and in Figure 5. We see that both LIM schemes provide an essential efficiency gain with respect to the EE scheme. On the N=512𝑁512N=512italic_N = 512 grid the gain of approximately a factor 10 is attained (see the bottom plot in Figure 5).

Table 8: Number of matrix-vector multiplications (matvecs), number of linear system solutions (for implicit schemes LSS and LIE), and achieved accuracy versus ττ\tauuproman_τ. ϵ=ϵ4(h)ϵsubscriptϵ4\epsilonup=\epsilonup_{4}(h)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h ), smoothed initial vector 𝐜0superscript𝐜0\mathbf{c}^{0}bold_c start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT, space grid N=128𝑁128N=128italic_N = 128.
ττ\tauuproman_τ scheme # matvecs /// error
# lin.syst. solutions ×103absentsuperscript103\times 10^{3}× 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (4.4)
1.0×1041.0superscript1041.0\times 10^{-4}1.0 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT LIM-LSS 34 / — 7.46×1027.46superscript1027.46\times 10^{-2}7.46 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT
LSS 2 / 2 1.37×1011.37superscript1011.37\times 10^{-1}1.37 × 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT
LIM-LIE 34 / — 6.06×1036.06superscript1036.06\times 10^{-3}6.06 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT
5.0×1065.0superscript1065.0\times 10^{-6}5.0 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT LIM-LSS 200 / — 3.42×1033.42superscript1033.42\times 10^{-3}3.42 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT
LSS 40 / 40 5.25×1035.25superscript1035.25\times 10^{-3}5.25 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT
LIM-LIE 185 / — 1.09×1041.09superscript1041.09\times 10^{-4}1.09 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
1.0×1061.0superscript1061.0\times 10^{-6}1.0 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT EE 200 / — 3.02×1053.02superscript1053.02\times 10^{-5}3.02 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
LIM-LSS 600 / — 8.14×1048.14superscript1048.14\times 10^{-4}8.14 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
LSS 200 / 200 1.04×1031.04superscript1031.04\times 10^{-3}1.04 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT
LIM-LIE 600 / — 1.50×1051.50superscript1051.50\times 10^{-5}1.50 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
Table 9: Number of matrix-vector multiplications (matvecs), number of linear system solutions (for implicit schemes LSS and LIE), and achieved accuracy versus ττ\tauuproman_τ. ϵ=ϵ4(h)ϵsubscriptϵ4\epsilonup=\epsilonup_{4}(h)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h ), smoothed initial vector 𝐜0superscript𝐜0\mathbf{c}^{0}bold_c start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT, space grid N=256𝑁256N=256italic_N = 256.
ττ\tauuproman_τ scheme # matvecs /// error
# lin.syst. solutions ×106absentsuperscript106\times 10^{6}× 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT (4.4)
1.0×1051.0superscript1051.0\times 10^{-5}1.0 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT LIM-LSS 0.38 / — 3.76×1033.76superscript1033.76\times 10^{-3}3.76 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT
LSS 0.02 / 0.02 9.55×1039.55superscript1039.55\times 10^{-3}9.55 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT
LIM-LIE 0.38 / — 2.27×1042.27superscript1042.27\times 10^{-4}2.27 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
5.0×1075.0superscript1075.0\times 10^{-7}5.0 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT LIM-LSS 2 / — 2.51×1042.51superscript1042.51\times 10^{-4}2.51 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
LSS 0.4 / 0.4 4.75×1044.75superscript1044.75\times 10^{-4}4.75 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
LIM-LIE 2 / — 1.99×1071.99superscript1071.99\times 10^{-7}1.99 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT
1.0×1071.0superscript1071.0\times 10^{-7}1.0 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT EE 2 / — 2.76×1062.76superscript1062.76\times 10^{-6}2.76 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT
LIM-LSS 6 / — 6.49×1056.49superscript1056.49\times 10^{-5}6.49 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
LSS 2 / 2 9.49×1059.49superscript1059.49\times 10^{-5}9.49 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
LIM-LIE 6 / — 9.59×1079.59superscript1079.59\times 10^{-7}9.59 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT
Refer to caption
Refer to caption
Figure 5: Achieved accuracy versus the number of matrix-vector multiplications (matvecs) for the EE scheme and the LIM schemes with the smoothed initial vector 𝐜0superscript𝐜0\mathbf{c}^{0}bold_c start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT and ϵ=ϵ4(1/64)ϵsubscriptϵ4164\epsilonup=\epsilonup_{4}(1/64)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( 1 / 64 ) on the N=256𝑁256N=256italic_N = 256 grid (upper plot) and the N=512𝑁512N=512italic_N = 512 grid (bottom plot). Increasing ττ\tauuproman_τ in the EE scheme further is impossible due to the stability restrictions.
Table 10: Number of matrix-vector multiplications (matvecs), number of linear system solutions (for implicit schemes LSS and LIE), and achieved accuracy versus ττ\tauuproman_τ. ϵ=ϵ4(h)ϵsubscriptϵ4\epsilonup=\epsilonup_{4}(h)roman_ϵ = roman_ϵ start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_h ), smoothed initial vector 𝐜0superscript𝐜0\mathbf{c}^{0}bold_c start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT, space grid N=512𝑁512N=512italic_N = 512.
ττ\tauuproman_τ scheme # matvecs /// error
# lin.syst. solutions ×106absentsuperscript106\times 10^{6}× 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT (4.4)
8.0×1078.0superscript1078.0\times 10^{-7}8.0 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT LIM-LSS 5.75 / — 3.14×1043.14superscript1043.14\times 10^{-4}3.14 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
LSS 0.25 / 0.25 7.13×1047.13superscript1047.13\times 10^{-4}7.13 × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT
LIM-LIE 5.75 / — 3.83×1063.83superscript1063.83\times 10^{-6}3.83 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT
4.0×1084.0superscript1084.0\times 10^{-8}4.0 × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT LIM-LSS 25 / — 1.65×1051.65superscript1051.65\times 10^{-5}1.65 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
LSS 5 / 5 3.56×1053.56superscript1053.56\times 10^{-5}3.56 × 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT
LIM-LIE 25 / — 9.87×1089.87superscript1089.87\times 10^{-8}9.87 × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT
8.0×1098.0superscript1098.0\times 10^{-9}8.0 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT EE 25 / — 1.98×1071.98superscript1071.98\times 10^{-7}1.98 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT
LIM-LSS 75 / — 4.44×1064.44superscript1064.44\times 10^{-6}4.44 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT
LSS 25 / 25 7.13×1067.13superscript1067.13\times 10^{-6}7.13 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT
LIM-LIE 75 / — 5.14×1085.14superscript1085.14\times 10^{-8}5.14 × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT

5.  Conclusions

The presented results allow to make the following conclusions.

  1. 1)

    Proposed local iteration schemes LIM-LSS and LIM-LIE (3.17), which are based on the implicit schemes LSS (3.14) and LIE (3.20), proved to be reliable in practice. They combine a simplicity and parallelism of explicit schemes with stability of implicit schemes. Theoretical estimates of efficiency for the LIM schemes are confirmed in the tests: the LIM schemes provide an efficiency gain up to a factor 10 with respect to the explicit scheme EE. The gain factor grows as the space grid gets finer.

  2. 2)

    In the considered numerical tests, gradient stable schemes based on the Eyre splitting appear to be less accurate than regular linearized schemes. In particular, the regular linearized implicit Euler scheme LIE (3.20) and based on it local iteration scheme LIM-LIE provide a higher accuracy than the linear stabilized splitting scheme LSS (3.14) and the local iteration scheme LIM-LSS. In the latter schemes a gradient stability is achieved by splitting, which leads to an additional error.

  3. 3)

    In cases where the diffusion boundary width ϵϵ\epsilonuproman_ϵ is chosen proportional to the grid size, i.e., ϵ=𝒪(h)ϵ𝒪\epsilonup=\mathcal{O}(h)roman_ϵ = caligraphic_O ( italic_h ), the time step size in the explicit scheme EE is bounded as τ=𝒪(h2)τ𝒪superscript2\tauup=\mathcal{O}(h^{2})roman_τ = caligraphic_O ( italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). Therefore, we can expect that the potential of the local iteration schemes for the Cahn-Hilliard equation is comparable to that for parabolic problems.

  4. 4)

    For non-smooth initial data, for instance, if the initial value vector is chosen randomly, additional accuracy restrictions on the time step size ττ\tauuproman_τ arise at times t𝒪(ϵ2)𝑡𝒪superscriptϵ2t\leqslant\mathcal{O}(\epsilonup^{2})italic_t ⩽ caligraphic_O ( roman_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). Since decreasing the time step size in an implicit scheme usually means lowering its efficiency, the local iteration schemes appear to be especially attractive (in these schemes decreasing the time step size does lead to lower computational costs). In such problems it seems sensible to apply local iterations schemes with an adaptive choice of the time step size [65].

Acknowledgments

The authors thank V.T. Zhukov (KIAM RAS) for useful discussions and consulting on local iterations schemes.

References

  • 1. Cahn, J.W., Hilliard, J.E. Free Energy of a Nonuniform System. I. Interfacial Free Energy // The Journal of Chemical Physics. 1958. V. 28. № 2. pp. 258-267. https://doi.org/10.1063/1.1744102
  • 2. Gurtin, M.E. Generalized Ginzburg-Landau and Cahn-Hilliard equations based on a microforce balance // Physica D: Nonlinear Phenomena.1996. V. 92. Iss. 3–4., pp. 178-192. https://doi.org/10.1016/0167-2789(95)00173-5
  • 3. Provatas, N., Elder, K. Phase-Field Methods in Materials Science and Engineering. First published:7 October 2010 DOI:10.1002/9783527631520. 2010 Wiley-VCH Verlag GmbH & Co. KGaA
  • 4. Steinbach, I., Salama, H. Lectures on Phase Field. Springer Cham, 2023. https://doi.org/10.1007/978-3-031-21171-3
  • 5. Skripov V P, Skripov A V ‘‘Spinodal decomposition (phase transitions via unstable states)’’ Sov. Phys. Usp. 22 389–410 (1979) https://doi.org/10.3367/UFNr.0128.197906a.0193
  • 6. Hohenberg. P.C., Halperin, B.I. Theory of dynamic critical phenomena // Rev. Mod. Phys. 177. V. 49. Iss. 3. pp. 435. https://doi.org/10.1103/RevModPhys.49.435
  • 7. Penrose, O., Fife, P.C. Thermodynamically consistent models of phase-field type for the kinetic of phase transitions // Physica D: Nonlinear Phenomena. 1990. V. 43. Iss. 1. pp. 44-62. https://doi.org/10.1016/0167-2789(90)90015-H.
  • 8. Bray, A.J. Theory of phase-ordering kinetics // Advances in Physics. 2002. V. 51, № 2. pp. 481-587. https://doi.org/10.1080/00018730110117433
  • 9. Miranville, A. The Cahn–Hilliard Equation: Recent Advances and Applications // Society for Industrial and Applied Mathematics. 2019. https://doi.org/10.1137/1.9781611975925
  • 10. Pego, R.L. Front Migration in the Nonlinear Cahn-Hilliard Equation // Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences. 1989. V. 422, No. 863. pp. 261-278. http://www.jstor.org/stable/2398477
  • 11. Bates, P.W.,Fife, P.C The Dynamics of Nucleation for the Cahn-Hilliard Equation // SIAM Journal on Applied Mathematics. 1993. V. 53. No. 4. pp. 990–1008. http://www.jstor.org/stable/2102259.
  • 12. de Mello, E.V.L., Otton Teixeira da Silveira Filho Numerical study of the Cahn–Hilliard equation in one, two and three dimensions // Physica A: Statistical Mechanics and its Applications. 2005. V. 347. pp. 429-443, https://doi.org/10.1016/j.physa.2004.08.076
  • 13. Vollmayr-Lee, B.P., Rutenberg, A.D. Fast and accurate coarsening simulation with an unconditionally stable time step // Phys. Rev. E. 2003. V. 68. Iss. 6. p. 066703. https://doi.org/10.1103/PhysRevE.68.066703
  • 14. Eyre, D.J. An unconditionally stable one-step scheme for gradient systems // Tech. report, Department of Mathematics, University of Utah. 1997. unpublished. https://api.semanticscholar.org/CorpusID:117273508
  • 15. Eyre, D.J. Unconditionally gradient stable time marching the Cahn-Hilliard equation // Computational and Mathematical Models of Microstructural Evolution, Mater. Res. Soc. Symp. Proc., ed., vol. 529, Bullard, J.W. and Chen, L.-Q. and Kalia, R.K. and Stoneham, A.M., 1998, pp. 39–46.
  • 16. Tierra, G., Guillén-González, F. Numerical methods for solving the Cahn-Hilliard equation and its applicability to related Energy-based models // Nečas Center for Mathematical Modeling, Preprint no. 2013-035.
  • 17. Cueto-Felgueroso, L., Peiraire, J. A time-adaptive finite volume method for the Cahn–Hilliard and Kuramoto–Sivashinsky equations // J. Comput. Phys. 2008. V. 227. Iss. 4. pp. 9985–10017. https://doi.org/10.1016/j.jcp.2008.07.024
  • 18. Li, Y., Choi, Y., Kim, J. Computationally efficient adaptive time step method for the Cahn–Hilliard equation // Computers & Mathematics with Applications. 2017. V. 73. Iss. 8. pp. 1855-1864. https://doi.org/10.1016/j.camwa.2017.02.021
  • 19. Zhang, Z., Qiao, Z. An Adaptive Time-Stepping Strategy for the Cahn-Hilliard Equation // Communications in Computational Physics. 2012. V. 11. Iss.4. pp. 1261-1278. https://doi.org/10.4208/cicp.300810.140411s
  • 20. Minkoff, S. E., Kridler, N. M. A comparison of adaptive time stepping methods for coupled flow and deformation modeling // Appl. Math. Model. 2006. V. 30. Iss. 9. pp. 993–1009. https://doi.org/10.1016/j.apm.2005.08.002
  • 21. Luo, F., Tang, T., Xie, H. Parameter-Free Time Adaptivity Based on Energy Evolution for the Cahn-Hilliard Equation // Communications in Computational Physics. 2016. V. 19,. Iss. 5. pp. 1542-1563. https://doi.org/10.4208/cicp.scpde14.45s
  • 22. Guillén-González, F., Tierra, G. Second order schemes and time-step adaptivity for Allen-Cahn and Cahn-Hilliard models // Computers and Mathematics with Applications. 2014. V. 68. Iss. 8. pp. 821–846. https://doi.org/10.1016/j.camwa.2014.07.014
  • 23. Kassam, A., Trefethen, L., Fourth-order time-stepping for stiff PDEs // SIAM J. Sci. Comput. 2005. V. 26. Iss. 4. pp. 1214–1233. https://doi.org/10.1137/S1064827502410633
  • 24. He, Y., Liu, Y., Tang, T. On large time-stepping methods for the Cahn–Hilliard equation // Applied Numerical Mathematics. 2007. V. 57. Iss. 5–7. pp. 616-628. https://doi.org/10.1016/j.apnum.2006.07.026
  • 25. Song, H. Energy stable and large time-stepping methods for the Cahn–Hilliard equation // International Journal of Computer Mathematics. 2015. V. 92. Iss. 10. pp. 2091-2108. https://doi.org/10.1080/00207160.2014.964694
  • 26. Li, D. Why large time-stepping methods for the Cahn-Hilliard equation is stable // Math. Comp. 2022. V. 91. № 238. pp. 2501-2515. https://doi.org/10.1090/mcom/3768
  • 27. Chen, W., Wang, C., Wang, X., Wise, S.M. Positivity-preserving, energy stable numerical schemes for the Cahn-Hilliard equation with logarithmic potential // Journal of Computational Physics: X. 2009. V. 3. pp. 100031, https://doi.org/10.1016/j.jcpx.2019.100031
  • 28. Chen, W., Wang, X., Yan, Y., Zhang, Z. A Second Order BDF Numerical Scheme with Variable Steps for the Cahn-Hilliard Equation // SIAM Journal on Numerical Analysis. 2019. V. 57. Iss. 1. pp. 495-525. https://doi.org/10.1137/18M1206084
  • 29. Zhang, J., Jiang, M., Gong, Y., & Zhao, J. Energy-stable predictor-corrector schemes for the Cahn-Hilliard equation // J. Comput. Appl. Math. 2020. V. 376. pp. 112832. https://doi.org/10.1016/j.cam.2020.112832
  • 30. Zhou, Q., Sun, Y. Energy stability of exponential time differencing schemes for the nonlocal Cahn-Hilliard equation // Numer. Methods Partial Differ. Eq. 2023. V. 39. Iss. 5. pp. 4030–4058. https://doi.org/10.1002/num.23035
  • 31. Lee, S. Unconditionally strong energy stable scheme for Cahn–Hilliard equation with second-order temporal accuracy // Math Meth Appl Sci. 2023. V. 46. Iss. 6. pp. 6463-6469. https://doi.org/10.1002/mma.8917
  • 32. Boyer, F., Minjeaud, S. Numerical schemes for a three component Cahn-Hilliard model // ESAIM: Mathematical Modelling and Numerical Analysis. 2011. V. 45. No. 4. pp. 697-738. https://doi.org/10.1051/m2an/2010072
  • 33. Brachet, M., Chehab, J.-P. Fast and Stable Schemes for Phase Fields Models // Computers & Mathematics with Applications. 2020. V. 80. Iss. 6. pp. 1683-1713. https://doi.org/10.1016/j.camwa.2020.07.015
  • 34. Elliott, C., French, D.A. A nonconforming finite element method for the two- dimensional Cahn–Hilliard equation // SIAM J. Numer. Anal. 1989. V. 26. No. 4. pp. 884–903. https://www.jstor.org/stable/2157884
  • 35. Barrett, J.B. An error bound for the finite element approximation of the Cahn–Hilliard equation with logarithmic free energy // Numer. Math. 1995. Vol. 72. pp. 1–20. https://doi.org/10.1007/s002110050157
  • 36. Chen, L.-Q., Shen, J., Applications of semi-implicit Fourier-spectral method to phase field equations // Comput. Phys. Commun. 1996. V. 108. Iss. 2-3. pp. 147–158. https://doi.org/10.1016/S0010-4655(97)00115-X
  • 37. Furihata, D. A stable and conservative finite difference scheme for the Cahn– Hilliard equation // Numer. Math. 2001. V. 87. Iss. 4. pp. 675–699. https://doi.org/10.1007/PL00005429
  • 38. Feng, X., Prohl, A. Error analysis of a mixed finite element method for the Cahn– Hilliard equation // Numer. Math. 2004.V. 99. Iss. 1. pp. 47–84. https://doi.org/10.1007/s00211-004-0546-5
  • 39. Wells, E., Kuhl, K., Garikipati A discontinuous Galerkin method for the Cahn– Hilliard equation // J. Comput. Phys. 2006. V. 218. Iss. 2. pp. 860–877. https://doi.org/10.1016/j.jcp.2006.03.010
  • 40. Wise, S.M., Wang, C., Lowengrub, J.S. An Energy-Stable and Convergent Finite-Difference Scheme for the Phase Field Crystal Equation // SIAM Journal on Numerical Analysis. 2009. V. 47. Iss. 3. pp. 2269-2288. https://doi.org/10.1137/0807381
  • 41. Du, Q., Ju, L., Tian, L. Finite element approximation of the Cahn–Hilliard equation on surfaces // Computer Methods in Applied Mechanics and Engineering. 2011. V. 200. Iss. 29–32. pp. 458-2470. https://doi.org/10.1016/j.cma.2011.04.018.
  • 42. Xia, Y., Xu, Y., Shu, C.-W. Local discontinuous Galerkin methods for the Cahn–Hilliard type equations // Journal of Computational Physics. 2007. V. 227. Iss. 1. pp. 472-491. https://doi.org/10.1016/j.jcp.2007.08.001
  • 43. Brenner, S.C., Diegel, A.E. Sung, L.-Y. A robust solver for a second order mixed finite element method for the Cahn–Hilliard equation // Journal of Computational and Applied Mathematics. 2020. V. 364. pp. 112322. https://doi.org/10.1016/j.cam.2019.06.038
  • 44. Gómez, H., Calo, V.M., Bazilevs, Y., Hughes, T.J.R Isogeometric analysis of the Cahn–Hilliard phase-field model // Computer Methods in Applied Mechanics and Engineering. 2008. V. 197, Iss. 49–50. pp. 4333-4352. https://doi.org/10.1016/j.cma.2008.05.003
  • 45. Zhang, R., Qian, X. Triangulation-based isogeometric analysis of the Cahn–Hilliard phase-field model // Computer Methods in Applied Mechanics and Engineering. 2019. V. 357. pp. 112569. https://doi.org/10.1016/j.cma.2019.112569
  • 46. Kästner, M., Metsch, P., de Borst, R. Isogeometric analysis of the Cahn–Hilliard equation –– a convergence study // Journal of Computational Physics. 2016. V. 305. pp. 360-371. https://doi.org/10.1016/j.jcp.2015.10.047
  • 47. Goudenège, L., Martin, D., Vial, G. High Order Finite Element Calculations for the Cahn-Hilliard Equation // J. Sci. Comput. 2012. V. 52. pp. 294–321 https://doi.org/10.1007/s10915-011-9546-7
  • 48. Čžao-Din, Y. On the stability of difference schemes for the solutions of parabolic differential equations. (Russian), Dokl. Akad. Nauk SSSR 117, 578–581 (1957) https://zbmath.org/?q=an:0102.33501
  • 49. Čžao-Din, Y. Some difference schemes for the numerical solution of differential equations of parabolic type.(Russian) Mat. Sb. (N.S.), 50(92), (1960), 391–422. https://www.mathnet.ru/eng/sm4800
  • 50. Gel’fand I.M., Lokutsievskii O.V. On difference schemes for solving the heat equation (Russian), In book: Godunov S.K., Ryaben’kii V.S., Introduction to the theory of difference schemes, Gosudarstv. Izdat. Fiz.-Mat. Lit., Moscow, 1962, 340 pp.
  • 51. Babenko, K. I. Foundations of numerical analysis (Russian), Publisher ‘‘Regular and chaotic dynamics’’, Izhevsk, 2002, 848 pp.
  • 52. Lokutsievskii V.O., Lokutsievskii O.V. Application of Chebyshev parameters to numerical solution of some evolution problems (Russian), KIAM Preprint № 99, Moscow, 1984, 30 pp. https://library.keldysh.ru/preprint.asp?id=1984-99
  • 53. Zhukov V.T. Numerical experiments for solving the heat equation by the local iteration method (Russian), KIAM Preprint № 97, Moscow, 1984, 22 pp. https://library.keldysh.ru/preprint.asp?id=1984-97
  • 54. Lokutsievskii, V.O., Lokutsievskii, O.V. On numerical solution of boundary value problems for equations of parabolic type. Sov. Math. Dokl. 34(3), 512–516 (1987) https://www.mathnet.ru/eng/dan47741
  • 55. Zhukov V.T. Local iteration difference schemes for parabolic equations (Russian), KIAM Preprint № 173, Moscow, 1986, 31 pp. https://library.keldysh.ru/preprint.asp?id=1986-173
  • 56. V.T. Zhukov Explicit Iteration Schemes for Parabolic Equations (Russian), Vopr. Atomn. Nauki I Tekn., Ser. Mat. Mod. Fiz. Proc., No. 3, pp. 40–46 (1993).
  • 57. Shvedov, A.S., Zhukov, V.T. Explicit iterative difference schemes for parabolic equations // Russian J. Numer. Anal. Math. Modelling. 1998. V. 13. № 2. pp. 133–148.
  • 58. Zhukov, V.T. Explicit methods of numerical integration for parabolic equations. Math. Models Comput. Simul., 3, pp. 311–332 (2011). https://doi.org/10.1134/S2070048211030136
  • 59. Zhukov V.T., Novikova N.D., Feodoritova O.B. On application of multigrid and explicit-iterative methods to solution of the parabolic equations with anisotropic discontinuous coefficients (Russian). KIAM Preprint № 85, Moscow, 2014, 24 pp. https://library.keldysh.ru/preprint.asp?lg=e&id=2014-85
  • 60. Zhukov V. T., Feodoritova O. B., Duben A.P., Novikova N.D. Explicit time integration of the Navier-Stokes equations using the local iteration method (Russian). KIAM Preprint № 12, Moscow, 2019, 32 pp. https://library.keldysh.ru/preprint.asp?lg=e&id=2019-12
  • 61. Zhukov, V.T., Feodoritova, O.B. On Development of Parallel Algorithms for Solving Parabolic and Elliptic Equations. J. Math. Sci. 254, 606–624 (2021). https://doi.org/10.1007/s10958-021-05329-y
  • 62. Zhukov, V.T., Zaitsev, N.A., Lysov, V.G. et al. Numerical analysis of a model of metal solidification, 2D case. Math Models Comput Simul 4, 440–453 (2012). https://doi.org/10.1134/S2070048212040096
  • 63. Lee, D., Huh, J.-Y., Jeong, D., Shin, J., Yun, A., Kim, J. Physical, mathematical, and numerical derivations of the Cahn–Hilliard equation // Computational Materials Science. 2014. V. 81. pp. 216-225. https://doi.org/10.1016/j.commatsci.2013.08.027
  • 64. Lee, S., Lee, C., Lee, H., Kim, J. Comparison of different numerical schemes for the Cahn-Hilliard equation // Journal of the Korea Society for Industrial and Applied Mathematics. 2013. V. 17. Iss. 3. pp. 197-207. https://doi.org/10.12941/jksiam.2013.17.197
  • 65. Botchev M.A., Zhukov V.T. Adaptive iterative explicit time integration for nonlinear heat conduction problems. Lobachevskii J Math 45, 12–20 (2024). https://doi.org/10.1134/S1995080224010086