We consider a time semidiscretization of the Ginzburg-Landau equation by the backward Euler scheme. For each time step τ, we build an exponential attractor of the dynamical system associated to the scheme. We prove that, as τ tends to 0, this attractor converges for the symmetric Hausdorff distance to an exponential attractor of the dynamical system associated to the Allen-Cahn equation. We also prove that the fractal dimension of the exponential attractor and of the global attractor is bounded by a constant independent of τ.
Citation: Narcisse Batangouna. A robust family of exponential attractors for a time semi-discretization of the Ginzburg-Landau equation[J]. AIMS Mathematics, 2022, 7(1): 1399-1415. doi: 10.3934/math.2022082
[1] | Xiao-Yu Li, Yu-Lan Wang, Zhi-Yuan Li . Numerical simulation for the fractional-in-space Ginzburg-Landau equation using Fourier spectral method. AIMS Mathematics, 2023, 8(1): 2407-2418. doi: 10.3934/math.2023124 |
[2] | Dieunel DOR . On the modified of the one-dimensional Cahn-Hilliard equation with a source term. AIMS Mathematics, 2022, 7(8): 14672-14695. doi: 10.3934/math.2022807 |
[3] | Aymard Christbert Nimi, Daniel Moukoko . Global attractor and exponential attractor for a Parabolic system of Cahn-Hilliard with a proliferation term. AIMS Mathematics, 2020, 5(2): 1383-1399. doi: 10.3934/math.2020095 |
[4] | Joseph L. Shomberg . Well-posedness and global attractors for a non-isothermal viscous relaxationof nonlocal Cahn-Hilliard equations. AIMS Mathematics, 2016, 1(2): 102-136. doi: 10.3934/Math.2016.2.102 |
[5] | Jianbo Yuan, Shixuan Zhang, Yongqin Xie, Jiangwei Zhang . Exponential attractors for nonclassical diffusion equations with arbitrary polynomial growth nonlinearity. AIMS Mathematics, 2021, 6(11): 11778-11795. doi: 10.3934/math.2021684 |
[6] | Binbin Miao, Chongbin Xu, Caidi Zhao . Statistical solution and piecewise Liouville theorem for the impulsive discrete Zakharov equations. AIMS Mathematics, 2022, 7(5): 9089-9116. doi: 10.3934/math.2022505 |
[7] | Chaeyoung Lee, Seokjun Ham, Youngjin Hwang, Soobin Kwak, Junseok Kim . An explicit fourth-order accurate compact method for the Allen-Cahn equation. AIMS Mathematics, 2024, 9(1): 735-762. doi: 10.3934/math.2024038 |
[8] | Tingting Liu, Tasneem Mustafa Hussain Sharfi, Qiaozhen Ma . Time-dependent asymptotic behavior of the solution for evolution equation with linear memory. AIMS Mathematics, 2023, 8(7): 16208-16227. doi: 10.3934/math.2023829 |
[9] | Junling Sun, Xuefeng Han . Existence of Sobolev regular solutions for the incompressible flow of liquid crystals in three dimensions. AIMS Mathematics, 2022, 7(9): 15759-15794. doi: 10.3934/math.2022863 |
[10] | Mourad S. Semary, M. T. M. Elbarawy, Aisha F. Fareed . Discrete Temimi-Ansari method for solving a class of stochastic nonlinear differential equations. AIMS Mathematics, 2022, 7(4): 5093-5105. doi: 10.3934/math.2022283 |
We consider a time semidiscretization of the Ginzburg-Landau equation by the backward Euler scheme. For each time step τ, we build an exponential attractor of the dynamical system associated to the scheme. We prove that, as τ tends to 0, this attractor converges for the symmetric Hausdorff distance to an exponential attractor of the dynamical system associated to the Allen-Cahn equation. We also prove that the fractal dimension of the exponential attractor and of the global attractor is bounded by a constant independent of τ.
We consider the following system: find u:Ω×R+⟶Rm(m⩾1) such that
∂u∂t−Δu=(1−|u|2)u,x∈Ω, t>0, | (1.1) |
where Ω is a bounded subset of Rd (d≥1) with smooth boundary ∂Ω. This system is endowed with homogeneous Dirichlet boundary conditions and an initial condition.
This problem arises in the study of superconductivity of liquids. The unknown u is an order parameter and when m=2 or 3, it can be interpreted as the preferential orientation vector of molecules (see, e.g., [4,11] and references therein). The set Ω is the region occupied by the liquid. We note that (1.1) is a system of reaction-diffusion equations. Indeed, by noting u=(u1,…,um), it can be written as
{∂u1∂t−Δu1=(1−∑mi=1u2i)u1,x∈Ω,t>0⋮∂um∂t−Δum=(1−∑mi=1u2i)um,x∈Ω,t>0. | (1.2) |
The boundary condition reads
ui=0 on x∈∂Ω, t>0,∀i∈{1,…,m}. |
When m=2, the system (1.1) is known as the Ginzburg-Landau equation. When m=1, the system reduces to a single equation called the Allen-Cahn equation [1].
Problem (1.1) has been extensively studied. In particular, starting with an initial value in L∞(Ω)m, it is easy to derive an L∞ bound on the solution and to obtain global existence. This problem illustrates the case of reaction-diffusion systems with an invariant region. In [18], Temam proved the existence of a global attractor associated to this problem. He also gave an upper bound for its Hausdorff dimension and for its fractal dimension, using the method of Lyapunov exponents. In this paper, we consider a time semidiscretization of problem (1.1) and we prove the convergence, as the time step goes to 0, of a family of exponential attractors associated to the discrete problem.
For a dissipative dynamical system, an exponential attractor is a compact positively invariant set which contains the global attractor, has finite fractal dimension and attracts exponentially the trajectories. In comparison with the global attractor, an exponential attractor is expected to be more robust to perturbations: global attractors are generally upper semicontinuous with respect to perturbations, but the lower semicontinuity can be proved only in some particular cases (see e.g. [2,13,18]). This includes perturbations which are obtained by time and/or space discretizations of the governing equations [17,20]. We note that, contrary to the global attractor, an exponential attractor is generally not unique.
The notion of inertial manifold, an exponential attractor which is also a manifold, was introduced in [9]. In relation with (1.1), we note that families of inertial manifolds which are robust wich respect to time and space discretization were built in [12] for the complex Ginzburg-Landau equation in one space dimension. However, all known constructions of inertial manifolds are based on a restrictive condition, the so-called spectral gap condition. As a consequence, the existence of inertial manifolds is not known for many physically important equations, such as the two-dimensional Navier-Stokes equations.
Eden et al. gave in [6] a construction of exponential attractions based on a "squeezing property". They proved the continuity of exponential attractors for classical Galerkin approximations, but only up to a time shift (see also [8,10]). In [7], Efendiev, Miranville and Zelik proposed a construction of exponential attractors based on a "smoothing property" and on an appropriate error estimate, where the continuity holds without time shift. Their construction, which is valid in Banach spaces, has been adapted to many situations, including singular perturbations. We refer the reader to the review [13] and the references therein for more details.
In [14], Pierre used the result of Efendiev, Miranville and Zelig to analyze the case where the perturbation is a time semidiscretization of the underlying equation, and when the time step goes to 0. He first proposed an abstract construction of a robust family of exponential attractors, and then he applied it to the backward Euler time semidiscretization of the Allen-Cahn equation with a polynomial nonlinearity. The abstract result in [14] was also applied in [3] to the case of a time-splitting discretization of the Caginalp phase-field system. The construction was adapted in [15] to a finite element space semidiscretization of the Allen-Cahn equation.
Our purpose in this note is to show that the abstract result in [14] can also be applied to the model problem (1.1), for every space dimension d and for every m. We use a backward Euler scheme for the time discretization. The analysis is comparable to the case of the Allen-Cahn equation (m=1) performed in [14], but the estimates are somewhat simpler here thanks to the L∞ estimates. Our paper is outlined as follows. We first derive in Section 2 the estimates for the continuous-in-time problem. Then, in Section 3, we derive their discrete-in-time counterparts and we give the error estimate between the discrete and countinuous problems. In the last section, we are in position to apply the abstract result. For every time step τ, we build an exponential attractor Mτ for the discrete-in-time problem and we show that Mτ converges for the symmetric Hausdorff distance to an exponential attractor M0 of the continuous problem, as τ tends to 0. We also show that the fractal dimension of Mτ is bounded by a constant independent of τ. As a corollary of our analysis, we obtain the upper semicontinuity of the global attractor Aτ as τ tends to 0. Since Aτ⊂Mτ, the fractal dimension of the global attractors is also bounded by a constant independent of τ.
Let H=L2(Ω)m be equipped with the usual norm
|v|20=m∑i=1∫Ωv2idx |
for v=(v1,…,vm) and the associated inner product (⋅,⋅)0. We set V=H10(Ω)m endowed with the norm
‖v‖2=m∑i=1∫Ω|∇vi|2dx. |
We recall the Poincaré inequality: there exists a constant c0=c0(Ω) such that
|v|0≤c0‖v‖,∀v∈V. | (2.1) |
We define the function g: Rm⟶ Rm by
g(w)=(|w|2−1)w,∀w∈ Rm. |
Using the coordinates g(w)=(g1(w),…,gm(w)), we have
gi(w)=gi(w1,…,wm)=(m∑j=1w2j−1)wi. |
The problem (1.1) with homogeneous Dirichlet condition and initial condition can be written as:
∂u∂t−Δu+g(u)=0,x∈Ω,t>0 | (2.2) |
u=0,x∈∂Ω,t>0 | (2.3) |
u(x,0)=u0(x),x∈Ω. | (2.4) |
Throughout this paper, we shall assume that δ≥1 and we denote
Bδ={w∈Rm,|w|≤δ} |
the ball of Rm centered at 0 and with radius δ. In order to derive L∞ estimates, the basic idea is that for w∈∂Bδ, we have |w|=δ and so the vector −g(w)=(1−δ2)w points inside Bδ. The set Bδ is an invariant region. Let
L2(Ω;Bδ)={v∈L2(Ω)m,v(x)∈Bδ for a.e.x∈Ω}. |
Note that L2(Ω;Bδ)⊂L∞(Ω)m and that L2(Ω;Bδ) is a closed convex subset of L2(Ω)m. More precisely, we have the following well-posedness result [18]:
Theorem 2.1. For all u0∈L2(Ω;Bδ), the problem (2.2)−(2.4) has a unique solution for all time, u(t)∈L2(Ω;Bδ),∀t, u(t)∈L2(0,T;H10),∀T>0, and the mapping S0(t):u0⟼u(t) is continuous in L2(Ω)m, ∀t≥0.
Furthermore, if u0∈H10(Ω)m∩L2(Ω;Bδ), then
u∈C([0,T];H10(Ω)m)∩L2(0,T;H2(Ω)),∀T>0. |
This result defines the semigroup S0(t):u0⟼u(t) on L2(Ω;Bδ).
We first derive the dissipative estimates. Multiplying (1.1) by u, integrating over Ω, and using the following result
g(u)u=(|u|2−1)|u|2=(|u|2−12)2−14, | (2.5) |
we obtain
12ddt|u|20+‖u‖2+∫Ω(|u|2−12)2dx≤14|Ω|, | (2.6) |
where |Ω|=∫Ω1dx. Using the Poincaré inequality (2.1), we obtain
ddt|u|20+2c20|u|20≤12|Ω| |
that is
ddt|u|20+c1|u|20≤c′1, |
with c1=2/c20 and c′1=|Ω|/2. Using Gronwall's lemma, we obtain
|u(t)|20≤|u0|20e−c1t+c′1c1(1−e−c1t). |
This yields:
Proposition 2.2 (Absorbing set in H). If |u0|0≤R, then
|u(t)|0≤ρ0,∀t≥t0(R), |
where ρ20=1+c′1c1 and t0(R)=1c1log(R2).
In the following, we set r>0. Integrating (2.6) from t to t+r yields
2∫t+rt‖u‖2ds≤|u(t)|20+rc′1,∀t≥0 |
If |u0|0≤R and t≥t0(R), then we have
2∫t+rt‖u‖2ds≤ρ20+rc′1. | (2.7) |
Using the fact that ∀w∈Bδ,
|g(w)|=|(|w|2−1)||w|≤(δ2+1)δ=Cδ, | (2.8) |
and multiplying (2.2) by −Δu, we obtain
12ddt‖u‖2+|Δu|20=∫Ωg(u)Δudx≤|Ω|2C2δ+12|Δu|20. |
Thus, we have
ddt‖u‖2≤|Ω|C2δ,∀t≥0. | (2.9) |
Using (2.7), (2.9) and the uniform Gronwall lemma [18], we find:
Proposition 2.3 (Absorbing set in V). If |u0|0≤R, then
‖u(t)‖≤ρ1,∀t≥t0(R)+r, |
where
ρ21=ρ20+rc′12r+r|Ω|C2δ |
is independent of R.
Next, we show that u is Hölder continuous in time. Multiplying (2.2) by ∂u∂t and integrating over Ω, we obtain
|dudt|20+12ddt‖u‖2+ddt∫Ω14(|u|2−1)2dx=0. |
Integrating over [0,t], we have
∫t0|dudt|20+12‖u(t)‖2+∫ΩG(u(t))dx≤12‖u0‖2+∫ΩG(u0)dx,∀t≥0, |
where
G(w)=14(|w|2−1)2 for w∈Rm. |
In particular if u0∈L2(Ω;Bδ)∩V, we have for all t1,t2≥0,
|u(t1)−u(t2)|20=|∫t2t1duds(s)ds|20≤|t2−t1|C(‖u0‖,δ). | (2.10) |
We consider two solutions u and ˉu of (2.2)-(2.4) with values in Bδ. Let w=u−ˉu, which satisfies
∂w∂t−Δw+g(u)−g(ˉu)=0. | (2.11) |
We multiply (2.11) by w and we integrate over Ω. We find
12ddt|w|20+‖∇w‖2=−(g(u)−g(ˉu),u−ˉu)0≤cδ|w|20. | (2.12) |
In the last line, we have used the Cauchy-Schwarz inequality and the mean inequality, which reads
∀ v,ˉv∈Bδ,|g(v)−g(ˉv)|0≤cδ|v−ˉv|0, | (2.13) |
with
cδ=supw∈Bδ‖Dg(w)‖<+∞. |
Using Gronwall's lemma, we deduce from (2.12) that
|w(t)|20+2∫t0‖w(s)‖2ds≤|w(0)|20exp(2cδt),t≥0. | (2.14) |
Next, we multiply (2.11) by ∂w∂t and we integrate over Ω. We obtain
|∂w∂t|20+12ddt‖w‖2=−(g(u)−g(ˉu),∂w∂t)0, |
which implies
|∂w∂t|20+12ddt‖w‖2≤cδ|w|0|∂w∂t|0≤|∂w∂t|20+c2δ4|w|20. |
Thus, we have
ddt‖w‖2≤c2δ2|w|20,∀t≥0. | (2.15) |
Multiplying (2.15) by t and adding ‖w‖2, we find
ddt(t‖w‖2)≤c2δ2t|w|20+‖w‖2. |
Thus
t‖w(t)‖2≤∫t0c2δ2t|w(s)|20ds+∫t0‖w(s)‖2ds,∀t≥0. |
Using (2.14), we obtain that ∀t∈[0,T]
t‖w(t)‖2≤c1(δ,T)|w(0)|20. | (2.16) |
This is a H-V smoothing property.
We use the backward Euler scheme. Let τ>0 be the time step. The scheme can be written as: let u0∈L2(Ω;Bδ), and for n=0,1,…, find un+1∈L2(Ω;Bδ)∩V which solves
un+1−unτ−Δun+1+g(un+1)=0inV′. | (3.1) |
We have the following well-posedness result.
Theorem 3.1. Assume that δt≤1/cδ. Then for all u∈L2(Ω;Bδ), there exists a unique v=vτ,u∈L2(Ω;Bδ)∩V such that
v−uτ−Δv+g(v)=0inV′ | (3.2) |
In addition, the mapping Sτ:u⟼vτ,u is Lipschitz continuous from L2(Ω;Bδ) into L2(Ω;Bδ)∩V, with
‖Sτu−Sτˉu‖≤c0τ|u−ˉu|0,∀u,ˉu∈L2(Ω;Bδ). | (3.3) |
This result defines for each τ∈(0,1/cδ] a discrete semigroup {Snτ,n∈N} on L2(Ω;Bδ). In the remainder of the manuscript, we will assume that τ∈(0,1/cδ].
Proof. We first prove the existence of a solution. We assume that d=1,2 or 3 so that H10(Ω)⊂L6(Ω). The general case can be obtained on replacing g with gδ, where for w∈Rm,
{gδ(w)=g(w) if |w|≤δ,gδ(w)=(δ2−1)w if |w|≥δ. | (3.4) |
Let u∈L2(Ω;Bδ). We minimize the functional
F(w)=12τ|w−u|20+12‖w‖2+∫ΩG(w(x))dx |
in V, and we get a solution v∈V of the Euler-Lagrange equation (3.2). Since H10(Ω)⊂L6(Ω), we have g(v)∈L2(Ω)m and by elliptic regularity,
v∈W2,2(Ω)m⊂C(ˉΩ)m. |
It remains to show that v(x)∈Bδ for all x∈Ω. Since Bδ is a closed convex subset of Rm, it is sufficient to show that for any hyperplane H containing Bδ, we have v(x)∈H, ∀x∈Ω. Let H={w∈Rm,⟨n,w⟩≤β} be such a hyperplane, where n=(n1,…,nm) is a vector of Rm of norm 1 and β≥δ. Here, ⟨n,w⟩=∑mi=1niwi is the usual inner product of Rm. The partial differential equation (3.2) can be written componentwise
vi−τΔvi=τ(1−|v|2)vi+ui |
in L2(Ω), for i=1,…,m. Multiplying the above equality by ni and summing over i, we have
ˉv−τΔˉv=τ(1−|v|2)ˉv+ˉu |
in L2(Ω), where ˉv=∑mi=1nivi and ˉu=∑mi=1niui. Since u∈L2(Ω;Bδ), we have ˉu(x)≤β a.e. x∈Ω and so
ˉv−β−τΔˉv≤τ(1−|v|2)ˉv a.e. in Ω. |
Multiplying by (ˉv−β)+, which belongs to H10(Ω), we have
∫Ω(ˉv−β)2++τ∫Ω|∇(ˉv−β)+|2≤∫Ωτ(1−|v|2)ˉv(ˉv−β)+≤0 |
since β≥δ≥1. Thus, ˉv(x)≤β, ∀x∈Ω as claimed.
Next, we prove the uniqueness of the solution. Let u,ˉu∈L2(Ω;Bδ) and let v,ˉv be the corresponding solutions of (3.2) in L2(Ω;Bδ)∩V. The difference w=v−ˉv satisfies
wτ−Δw+g(v)−g(ˉv)=u−ˉuτ. |
Multiplying the above equality by w and using (2.13), we get
1τ|w|20+‖w‖2−cδ|w|20≤1τ|u−ˉu|0|w|0. |
Using (2.1) and dividing by |w|0, we find
(1τ−cδ)|w|0+1c0‖w‖≤1τ|u−ˉu|0. |
If τ≤1/cδ, we find (3.3), which implies the uniqueness of the solution.
Proposition 3.2 (Absorbing set in H). Assume that τ≤c20/2. If |u0|0≤R, then
|u0|0≤ρ0,∀nτ≥2t0(R), |
where ρ0 and t0(R) are as in Proposition 2.2.
Proof. We multiply (3.1) by un+1 in H, we use (2.5) and the identity
(a−b,a)0=12(|a|20−|b|20+|a−b|20). | (3.5) |
This yields
12τ(|un+1|20−|un|20+|un+1−un|20)+‖un+1‖2≤|Ω|4. | (3.6) |
From the Poincaré inequality (2.1), we deduce that
12τ(|un+1|20−|un|20)+1c20|un+1|20≤|Ω|4, |
that is
(1+2τc20)|un+1|20≤|un|20+c′1τ, | (3.7) |
where c′1=|Ω|/2. We set α=1+2τc20. By induction, we deduce from (3.7) that
|un|20≤α−n|u0|20+c′1τ1−α−nα−1,∀n≥0. |
Using the inequality 1+x≥exp(x/2), valid for all x∈[0,1], we see that if 2τ≤c20, then we have α−1≤exp(−τ/c20). Thus, we get
|un|2≤exp(−nτ/c20)|u0|2+c′1c202,∀n≥0. |
The claim follows, by setting ρ20=1+c′1c202, as above.
We recall the following discrete uniform Gronwall lemma from [16].
Lemma 3.3. Let n0,n∈N and let a1, a2, a3, τ be positive constants. Assume that (dn), (gn) and (hn) are three sequences of nonnegative real numbers which satisfy
dn+1−dnτ≤gndn+hn,∀n≥n0, |
and
τk0+N∑n=k0gn≤a1,τk0+N∑n=k0hn≤a2,τk0+N∑n=k0dn≤a3, |
for all k0≥n0. Then
dn≤(a2+a3r′)exp(a1),∀n≥n0+N, |
where r′=τN.
We have:
Proposition 3.4 (Absorbing set in V). We assume that τ>0 satisfies τ≤c20/2 and τ≤r/2. If |u0|0≤R, then for all n∈N such that nτ≥2t0(R)+2r, we have
‖un‖0≤2ρ1. | (3.8) |
Proof. First, we multiply (3.1) by −Δun+1 and we integrate over Ω. This yields
12τ(‖un+1‖2−‖un‖2+‖un+1−un‖2)+|Δun+1|20=∫Ωg(u)Δu, |
where ∫Ωg(u)Δu=∫Ω∑mi=1gi(u)Δui. By the Cauchy-Schwarz inequality, we have
12τ(‖un+1‖2−‖un‖2+‖un+1−un‖2)+|Δun+1|20≤∫Ω|g(un+1)||Δun+1|. |
Since |un+1|≤δ for a.e. x∈Ω, we have |g(un+1)|≤Cδ (cf. (2.8)). Young's inequality yields
12τ(‖un+1‖2−‖un‖2+‖un+1−un‖2)+|Δun+1|20≤12|Ω|C2δ+12|Δun+1|20. |
Thus, we have
1τ(‖un+1‖2−‖un‖2)+1τ‖un+1−un‖2≤|Ω|C2δ,∀n>0. | (3.9) |
Next, let k0,n∈N∖{0}. By summing inequality (3.6) for n varying from k0−1 to k0+N−1, we find
|uk0+N|20−|uk0−1|20+2τk0+N−1∑n=k0−1‖un+1‖2≤|Ω|2τ(N+1). |
If k0τ≥2t0(R)+τ, we deduce from Proposition 3.2 that
2τk0+N∑n=k0‖un‖2≤|Ω|2τ(N+1)+ρ20. | (3.10) |
Let n0∈N such that n0τ≥2t0(R)+τ and let N=[r/τ]. We assume that τ≤r/2 so that N≥2. We set r′=Nτ∈[r−τ,r] and
k1=|Ω|(r′+r)/4+ρ20/2. |
We apply Lemma 3.3 with dn=‖un‖2, gn=0 and hn=|Ω|C2δ. Using the estimates (3.9) and (3.10), we obtain
‖un‖2≤k1r′+|Ω|C2δ(r′+r),∀n≥n0+N. |
Since r′∈[r/2,r], this implies the inequality (3.8).
In this section, (un) and (ˉun) are two sequences generated by the scheme (3.1). We denote by vn=un−ˉun their difference, which satisfies
1τ(vn+1−vn)−Δvn+1+g(un+1)−g(ˉun+1)=0 in V′. | (3.11) |
We first derive a discrete version of estimate 2.14.
Lemma 3.5. We assume that τ≤1/4cδ. Then for all n≥0, we have
|vn|20+2τn−1∑k=0‖vk+1‖2≤exp(4cδnτ)|v0|20. |
Proof. We multiply Eq (3.11) by vn+1, we integrate over Ω, we use the Cauchy-Scwharz inequality and estimate (2.13). This yields
12τ(|vn+1|20−|vn|20)+‖vn+1‖2≤cδ|vn+1|20, |
that is
(1−2τcδ)|vn+1|20+2τ‖vn+1‖2≤|vn|20, |
for all n≥0. Since
1≤11−s≤1+2s,∀s∈[0,12], |
it follows that
|vn+1|20+2τ‖vn+1‖2≤(1+4τcδ)|vn|20,∀n≥0. |
By induction, we obtain
|vn|20+2τn−1∑k=0‖vk+1‖2≤(1+4τcδ)n|v0|20,∀n≥0. |
Finally, we use that 1+s≤exp(s) with s=4τcδ and the claim is proved.
Next, we derive an H-V smoothing property, a discrete analog of (2.16).
Lemma 3.6. Let T>0. For all 0<nτ≤T, we have
nτ‖vn‖2≤C(T,δ)|v0|20. |
Proof. Let n≥1. We know by Theorem (3.1) that vn and vn+1 both belong to V. We multiply (3.11) by vn+1−vn and integrate over Ω. Using the identity (3.5) (with the norm |⋅|0 replaced by ‖⋅‖), we find
1τ|vn+1−vn|20+12‖vn+1‖2−12‖vn‖2+12‖vn+1−vn‖2=−(g(un+1)−g(ˉun+1),vn+1−vn)0≤cδ|vn+1|0|vn+1−vn|0≤1τ|vn+1−vn|20+c2δτ4|vn|20. |
In the second line, we used the Cauchy-Schwarz inequality and the estimate (2.13). Thus,
‖vn+1‖2−‖vn‖2≤c2δτ2|vn|20,∀n≥1. |
We multiply this by n and we add ‖vn+1‖2. This yields
(n+1)‖vn+1‖2−n‖vn‖2≤c2δ2nτ|vn|20+‖vn+1‖2,∀n≥1. |
We set αn=n‖vn‖2 (αn=0 if n=0) and βn=c2δ2nτ|vn|20+‖vn+1‖2. From what precedes, we have (note that the case n=0 is obvious)
αn+1≤αn+βn,∀n≥0. |
By induction, αn≤α0+∑n−1k=0βk, for all n≥1. Thus,
nτ‖vn‖2≤c2δ2τn−1∑k=0kτ|vk|20+τn−1∑k=0‖vk+1‖2. |
Applying Lemma 3.5, we find that for all n≥1,
nτ‖vn‖2≤c2δ2(nτ)2exp(4cδnτ)|v0|20+12exp(4cδnτ)|v0|20. |
Thus, if 0<nτ≤T, we may choose
C(T,δ)=c2δ2T2exp(4cδT)+12exp(4cδT), |
and this concludes the proof.
Proposition 3.7. Assume that τ≤1/4cδ. For each T>0 and each R>0, there is a constant C(T,R) independent of τ such that |u0|0≤R and 0≤nτ≤T imply
|un|0≤C(T,R). |
Proof. We set un=(un−ˉun)+(ˉun−ˉu0), where (ˉun) is the solution of the scheme (3.1) with initial value ˉu0=0. We write ˉun−ˉu0=∑n−1k=0(ˉuk+1−ˉuk) and by the triangle inequality, we have
|un|0≤|un−ˉun|0+n−1∑k=0|ˉuk+1−ˉuk|0. |
For the first term in the right-hand side, we use Lemma 3.5 with |v0|0≤R, and for the second term, we apply the Cauchy-Schwarz inequality. This yields
|un|0≤exp(2cδnτ)R+(nτ)1/2(1τn−1∑k=0|ˉuk+1−ˉuk|0)1/2, | (3.12) |
for all n≥0. Now, we estimate the last term above. We write (3.1) for the sequence (ˉun) and we multiply this by ˉun+1−ˉun for the L2(Ω) scalar product. This yields
|ˉun+1−ˉun|20τ+12‖ˉun+1‖2−12‖ˉun‖2≤|(g(ˉun+1),ˉun+1−ˉun)0|≤Cδ|Ω|1/2|ˉun+1−ˉun|0≤12τ|ˉun+1−ˉun|20+C2δτ2|Ω|, |
where Cδ is the constant from (2.8). By summing these estimates, we obtain
12τn−1∑k=0|ˉuk+1−ˉuk|20+12‖ˉun‖2≤12‖ˉu0‖2+C2δ2nτ|Ω|, |
for all n≥0. Since ˉu0=0, we find that
1τn−1∑k=0|ˉuk+1−ˉuk|20≤C2δnτ|Ω|. |
Using (3.12), we see that if 0≤nτ≤T, we have |un|0≤C(R,T) with
C(R,T)=exp(2cδT)R+CδT|Ω|1/2. |
This concludes the proof.
To estimate the error over a finite time interval, we follow the methodology described in [19]. We consider a sequence (un) generated by (3.1). For each τ>0, we associate to this sequence the functions uτ,ˉuτ:R+⟶L2(Ω) defined by
uτ(t)=un+t−nττ(un+1−un),t∈[nτ,(n+1)τ),ˉuτ(t)=un+1,t∈[nτ,(n+1)τ). |
We assume that u0∈L2(Ω;Bδ)∩V. By Theorem 3.1, for each n, un belongs to L2(Ω;Bδ)∩V. Thus, uτ∈C0(R+;L2(Ω;Bδ)∩V), ∂tuτ∈L∞loc(R+;V) and ˉuτ∈L∞loc(R+;L2(Ω;Bδ)∩V). The scheme (3.1) can be rewritten
∂tuτ−Δˉuτ+g(ˉuτ)=0 in V′, for a.e. t≥0, |
or in the same way,
∂tuτ−Δuτ+g(uτ)=−Δ(uτ−ˉuτ)+[g(uτ)−g(ˉuτ)], for a.e. t≥0. | (3.13) |
Let u be solution of (2.2)-(2.4) with u0∈L2(Ω;Bδ)∩V. We define eτ=uτ−u. The following error estimate holds:
Theorem 3.8. Let T>0 and R1>0. There exists a constant C(T,R1) independent of τ such that u0=u0 and ‖u0‖≤R1 imply
supt∈[0,Nτ]|eτ(t)|0≤C(T,R1)τ1/2, |
where N=⌊T/τ⌋ (here, ⌊⋅⌋ denotes the floor function).
Proof. By subtracting (2.2) from (3.13), we find
∂teτ−Δeτ+g(uτ)−g(u)=−Δ(uτ−ˉuτ)+[g(uτ)−g(ˉuτ)], for a.e. t≥0. |
Multiplying by eτ, and integrating over Ω, we obtain
12ddt|eτ|20+‖eτ‖2≤|(g(uτ)−g(u),eτ)0|+‖uτ−ˉuτ‖‖eτ‖+|(g(uτ)−g(ˉuτ),eτ)0|≤cδ|eτ|20+‖uτ−ˉuτ‖‖eτ‖+cδ|uτ−ˉuτ|0|eτ|0, | (3.14) |
where cδ is the constant from (2.13). From the Poincaré inequality (2.1) and Young's inequality, we derive that
ddt|eτ|20≤(2cδ+c2δc20)|eτ|20+‖uτ−ˉuτ‖2, for a.e. t≥0. |
Let T>0 and define N=⌊T/τ⌋. Using eτ(0)=0, the classic Gronwall lemma yields
|eτ(t)|20≤exp((2cδ+c20c2δ)T)∫Nτ0‖uτ−ˉuτ‖2ds,∀t∈[0,Nτ]. |
On [nτ,(n+1)τ], we have ‖uτ−ˉuτ‖≤‖un+1−un‖, so that
∫Nτ0‖uτ−¯uτ‖2Vds≤τN−1∑n=0‖un+1−un‖20. |
By summing estimate (3.9) from n=0 to n=N−1 (we note that (3.9) is also valid for n=0 since u0∈V), we find that
N−1∑n=0‖un+1−un‖2≤‖u0‖2+Nτ|Ω|C2δ. |
Hence,
|eτ(t)|20≤exp((2cδ+c20c2δ)T)(R21+T|Ω|C2δ)τ,∀t∈[0,Nτ]. |
This proves the claim.
We recall some standard definitions (see e.g. [13,18]). Throughout Section 4.1, K denotes a closed bounded subset of the Hilbert space H. A continuous-in-time semigroup {S(t), t∈R+} on K is a family of (nonlinear) operators such that S(t) is a continuous operator from K into itself, for all t≥0, with S(0)=Id (identity in K) and
S(t+s)=S(t)∘S(s),∀s,t≥0. |
A discrete-in-time semigroup {S(t), t∈N} on K is a family of (nonlinear) operators which satisfy these properties with R+ replaced by N. A discrete-in-time semigroup is usually denoted {Sn, n∈N}, where S(=S(1)) is a continuous (nonlinear) operator from K into itself. The term "dynamical system" will sometimes be used instead of "semigroup".
Definition 4.1 (Global attractor). Let {S(t), t≥0} be a continuous or discrete semigroup on K. A set A⊂K is called the global attractor of the dynamical system if the following three conditions are satisfied:
1. A is compact in H;
2. A is invariant, i.e. S(t)A=A, for all t≥0;
3. A attracts K, i.e.
limt→+∞distH(S(t)K,A)=0. |
Here, distH denotes the non-symmetric Hausdorff semidistance in H between two subsets, which is defined as
distH(A,B)=supa∈Ainfb∈B|a−b|H. |
It is easy to see, thanks to the invariance and the attracting property, that the global attractor, when it exists, is unique [18].
Let A⊂H be a (relatively compact) subset of H. For ϵ>0, we denote Nϵ(A,H) the minimum number of balls of H of radius ϵ>0 which are necessary to cover A. The fractal dimension of A (see e.g. [6,18]) is the number
dimF(A,H)=lim supϵ→0log2(Nϵ(A,H))log2(1/ϵ)∈[0,+∞]. |
Definition 4.2. (Exponential attractor). Let {S(t), t≥0} be a continuous or discrete semigroup on K. A set M⊂K is an exponential attractor of the dynamical system if the following three conditions are satisfied:
1. M is compact in H and has finite fractal dimension;
2. M is positively invariant, i.e. S(t)M⊂M, for all t≥0;
3. M attracts K exponentially, i.e.
distH(S(t)K,M)≤Ce−αt,t≥0, |
for some positive constants C and α.
The exponential attractor, if it exists, contains the global attractor (actually, the existence of an exponential attractor yields the existence of the global attractor, see [2,5]).
We may now state our main result. We recall that Kδ=L2(Ω;Bδ) is a closed convex subset of H and that V=H10(Ω)m is compactly imbedded into H. We also note that Kδ is a bounded subset of H since for each v∈Kδ, we have
|v|0≤δ|Ω|1/2. |
Theorem 2.1 shows that {S0(t) t∈R+} is a continuous-in-time semigroup on Kδ and Theorem 3.1 shows that {Snτ n∈N} is a discrete-in-time semigroup on Kδ. We have:
Theorem 4.3. Let τ0>0 be small enough. For every τ∈(0,τ0], {Snτ n∈N} possesses an exponential attractor Mτ on Kδ and {S0(t) t∈N+} possesses an exponential attractor M0 on Kδ such that:
1. The fractal dimension of Mτ is bounded, uniformly with respect to τ∈[0,τ0],
dimFMτ≤C1, |
where C1 is independent of τ;
2. Mτ attracts Kδ uniformly with respect to τ∈(0,τ], i.e.
∀τ∈(0,τ0], distH(SnτKδ,Mτ)≤C2e−C3nτ, n∈N, |
where the positive constants C2 and C3 are independent of τ;
3. the family {Mττ∈[0,τ0]} is continuous at 0,
distsym(Mτ,M0)≤C4τC5, |
where C4>0 and C5∈(0,1) are independent of τ and distsym is the symmetric Hausdorff distance between subsets of H, defined by
distsym(A,B)=max(distH(A,B),distH(B,A)). |
Proof. We apply Theorem 2 in [14] with the spaces H and V and the set
B={v∈Kδ : ‖v‖≤2ρ1}, |
and we choose τ0>0 small enough so that all the estimates from Section 3 are valid. By Proposition 2.3 and Proposition 3.4, B is an absorbing set in Kα, uniformly with respect to τ∈[0,τ0]. The estimates derived for the continuous problem in Section 2 show that assumptions (H1)-(H4) from [14, Theorem 2] are satisfied. The estimates from Section 3 show that assumptions (H5)-(H9) are also satisfied. The conclusion follows (we note that Theorem 2 in [14] is stated for a family of semigroups which act on the whole space H, but with a minor modification of the proof, it can be applied to our situation where the semigroups act on Kδ).
By arguing as in the proof of Corollary 1 in [14], we have:
Proposition 4.4. For each τ∈[0,τ0], the semigroup {Sτ(t),t≥0} possesses a global attractor Aτ in Kδ which is bounded in V, compact and connected in H. In addition, distH(Aτ,A0)→0 when τ→0+ and the fractal dimension of Aτ is bounded by a constant independent of τ.
The author declares no conflict of interest.
[1] |
S. Allen, J. Cahn, A microscopic theory for antiphase boundary motion and its application to antiphase domain coarsing, Acta Metall., 27 (1979), 1085–1095. doi: 10.1016/0001-6160(79)90196-2. doi: 10.1016/0001-6160(79)90196-2
![]() |
[2] | A. V. Babin, M. I. Vishik, Attractors of evolution equations, vol. 25 of Studies in Mathematics and its Applications, North-Holland Publishing Co., Amsterdam, 1992. doi: 10.1016/s0168-2024(08)x7020-1. |
[3] |
N. Batangouna, M. Pierre, Convergence of exponential attractors for a time splitting approximation of the {C}aginalp phase-field system, Commun. Pure Appl. Anal., 17 (2018), 1–19. doi: 10.3934/cpaa.2018001. doi: 10.3934/cpaa.2018001
![]() |
[4] |
C. Cavaterra, E. Rocca, H. Wu, Optimal boundary control of a simplified Ericksen-Leslie system for nematic liquid crystal flows in 2D, Arch. Ration. Mech. Anal., 224 (2017), 1037–1086. doi: 10.1007/s00205-017-1095-2. doi: 10.1007/s00205-017-1095-2
![]() |
[5] | V. V. Chepyzhov, M. I. Vishik, Attractors for equations of mathematical physics, vol. 49 of American Mathematical Society Colloquium Publications, American Mathematical Society, Providence, RI, 2002. doi: 10.1090/coll/049. |
[6] | A. Eden, C. Foias, B. Nicolaenko, R. Temam, Exponential attractors for dissipative evolution equations, vol. 37 of RAM: Research in Applied Mathematics, Masson, Paris; John Wiley & Sons, Ltd., Chichester, 1994. |
[7] |
M. Efendiev, A. Miranville, S. Zelik, Exponential attractors for a singularly perturbed Cahn-Hilliard system, Math. Nachr., 272 (2004), 11–31. doi: 10.1002/mana.200310186. doi: 10.1002/mana.200310186
![]() |
[8] |
P. Fabrie, C. Galusinski, A. Miranville, Uniform inertial sets for damped wave equations, Discrete Contin. Dynam. Systems, 6 (2000), 393–418. doi: 10.3934/dcds.2000.6.393. doi: 10.3934/dcds.2000.6.393
![]() |
[9] |
C. Foias, G. R. Sell, R. Temam, Inertial manifolds for nonlinear evolutionary equations, J. Differential Equations, 73 (1988), 309–353. doi: 10.1016/0022-0396(88)90110-6. doi: 10.1016/0022-0396(88)90110-6
![]() |
[10] | C. Galusinski, Perturbations singulières de problèmes dissipatifs : étude dynamique via l'existence et la continuité d'attracteurs exponentiels, PhD thesis, Université de Bordeaux, 1996. |
[11] |
F. Guillén-González, M. Samsidy Goudiaby, Stability and convergence at infinite time of several fully discrete schemes for a Ginzburg-Landau model for nematic liquid crystal flows, Discrete Contin. Dyn. Syst., 32 (2012), 4229–4246. doi: 10.3934/dcds.2012.32.4229. doi: 10.3934/dcds.2012.32.4229
![]() |
[12] |
G. J. Lord, Attractors and inertial manifolds for finite-difference approximations of the complex Ginzburg-Landau equation, SIAM J. Numer. Anal., 34 (1997), 1483–1512. doi: 10.1137/S003614299528554X. doi: 10.1137/S003614299528554X
![]() |
[13] | A. Miranville, S. Zelik, Attractors for dissipative partial differential equations in bounded and unbounded domains, in Handbook of differential equations: evolutionary equations. {V}ol. IV, Handb. Differ. Equ., Elsevier/North-Holland, Amsterdam, 4 (2008), 103–200. doi: 10.1016/S1874-5717(08)00003-0. |
[14] |
M. Pierre, Convergence of exponential attractors for a time semi-discrete reaction-diffusion equation, Numer. Math., 139 (2018), 121–153. doi: 10.1007/s00211-017-0937-z. doi: 10.1007/s00211-017-0937-z
![]() |
[15] |
M. Pierre, Convergence of exponential attractors for a finite element approximation of the Allen-Cahn equation, Numer. Funct. Anal. Optim., 39 (2018), 1755–1784. doi: 10.1080/01630563.2018.1497651. doi: 10.1080/01630563.2018.1497651
![]() |
[16] |
J. Shen, Long time stability and convergence for fully discrete nonlinear Galerkin methods, Appl. Anal., 38 (1990), 201–229. doi: 10.1080/00036819008839963. doi: 10.1080/00036819008839963
![]() |
[17] | A. M. Stuart, A. R. Humphries, Dynamical systems and numerical analysis, vol. 2 of Cambridge Monographs on Applied and Computational Mathematics, Cambridge University Press, Cambridge, 1996. |
[18] | R. Temam, Infinite-dimensional dynamical systems in mechanics and physics, vol. 68 of Applied Mathematical Sciences, Springer-Verlag, New York, second ed., 1997. doi: 10.1007/978-1-4612-0645-3. |
[19] |
X. Wang, Approximation of stationary statistical properties of dissipative dynamical systems: time discretization, Math. Comp., 79 (2010), 259–280. doi: 10.1090/S0025-5718-09-02256-X. doi: 10.1090/S0025-5718-09-02256-X
![]() |
[20] |
X. Wang, Numerical algorithms for stationary statistical properties of dissipative dynamical systems, Discrete Contin. Dyn. Syst., 36 (2016), 4599–4618. doi: 10.3934/dcds.2016.36.4599. doi: 10.3934/dcds.2016.36.4599
![]() |
1. | Dieunel Dor, Morgan Pierre, A robust family of exponential attractors for a linear time discretization of the Cahn-Hilliard equation with a source term, 2024, 58, 2822-7840, 1755, 10.1051/m2an/2024061 | |
2. | Dongsun Lee, Computing the area-minimizing surface by the Allen-Cahn equation with the fixed boundary, 2023, 8, 2473-6988, 23352, 10.3934/math.20231187 |