
This paper aimed to further introduce the finite element analysis of non-smooth data for semilinear stochastic subdiffusion problems driven by fractionally integrated additive noise. The mild solution of this stochastic model consisted of three different Mittag-Leffler functions. We analyzed the smoothness of the solution and utilized complex integration to approximate the error of the solution operator under non-smooth data. Consequently, optimal convergence estimates were obtained, and we also obtained the continuity conditions of the mild solution. Finally, the influence of the fractional parameters α and γ on the convergence rates were accurately demonstrated through numerical examples.
Citation: Fang Cheng, Ye Hu, Mati ur Rahman. Analyzing the continuity of the mild solution in finite element analysis of semilinear stochastic subdiffusion problems[J]. AIMS Mathematics, 2024, 9(4): 9364-9379. doi: 10.3934/math.2024456
[1] | Muhammad Imran Liaqat, Fahim Ud Din, Wedad Albalawi, Kottakkaran Sooppy Nisar, Abdel-Haleem Abdel-Aty . Analysis of stochastic delay differential equations in the framework of conformable fractional derivatives. AIMS Mathematics, 2024, 9(5): 11194-11211. doi: 10.3934/math.2024549 |
[2] | Xiaodong Zhang, Junfeng Liu . Solving a class of high-order fractional stochastic heat equations with fractional noise. AIMS Mathematics, 2022, 7(6): 10625-10650. doi: 10.3934/math.2022593 |
[3] | Snezhana Hristova, Antonia Dobreva . Existence, continuous dependence and finite time stability for Riemann-Liouville fractional differential equations with a constant delay. AIMS Mathematics, 2020, 5(4): 3809-3824. doi: 10.3934/math.2020247 |
[4] | Wedad Albalawi, Muhammad Imran Liaqat, Fahim Ud Din, Kottakkaran Sooppy Nisar, Abdel-Haleem Abdel-Aty . Well-posedness and Ulam-Hyers stability results of solutions to pantograph fractional stochastic differential equations in the sense of conformable derivatives. AIMS Mathematics, 2024, 9(5): 12375-12398. doi: 10.3934/math.2024605 |
[5] | Ramkumar Kasinathan, Ravikumar Kasinathan, Dumitru Baleanu, Anguraj Annamalai . Well posedness of second-order impulsive fractional neutral stochastic differential equations. AIMS Mathematics, 2021, 6(9): 9222-9235. doi: 10.3934/math.2021536 |
[6] | Zuliang Lu, Fei Cai, Ruixiang Xu, Lu Xing . Convergence and proposed optimality of adaptive finite element methods for nonlinear optimal control problems. AIMS Mathematics, 2022, 7(11): 19664-19695. doi: 10.3934/math.20221079 |
[7] | Lakhlifa Sadek, Tania A Lazǎr, Ishak Hashim . Conformable finite element method for conformable fractional partial differential equations. AIMS Mathematics, 2023, 8(12): 28858-28877. doi: 10.3934/math.20231479 |
[8] | Zhengang Zhao, Yunying Zheng, Xianglin Zeng . Finite element approximation of fractional hyperbolic integro-differential equation. AIMS Mathematics, 2022, 7(8): 15348-15369. doi: 10.3934/math.2022841 |
[9] | Feliz Minhós, Ali Raza, Umar Shafique . An efficient computational analysis for stochastic fractional heroin model with artificial decay term. AIMS Mathematics, 2025, 10(3): 6102-6127. doi: 10.3934/math.2025278 |
[10] | Changling Xu, Hongbo Chen . A two-grid $ P_0^2 $-$ P_1 $ mixed finite element scheme for semilinear elliptic optimal control problems. AIMS Mathematics, 2022, 7(4): 6153-6172. doi: 10.3934/math.2022342 |
This paper aimed to further introduce the finite element analysis of non-smooth data for semilinear stochastic subdiffusion problems driven by fractionally integrated additive noise. The mild solution of this stochastic model consisted of three different Mittag-Leffler functions. We analyzed the smoothness of the solution and utilized complex integration to approximate the error of the solution operator under non-smooth data. Consequently, optimal convergence estimates were obtained, and we also obtained the continuity conditions of the mild solution. Finally, the influence of the fractional parameters α and γ on the convergence rates were accurately demonstrated through numerical examples.
Fractional calculus has gained increasing attention due to its potential applications in various fields of science and engineering [19]. However, the study of fractional calculus has been predominantly focused on deterministic equations, using either deterministic or probabilistic methods. This approach limits the modeling of real-world phenomena where the propagation speed can be finite, as heat flow can be disrupted by material response. In contrast, the classical heat equation assumes infinite speed of heat flow. Recent studies have shown that materials with thermal memory can exhibit finite heat flow speed [2]. This is due to the convolution term in the definition of fractional derivatives and integrals, which implies that the nearer past has a stronger influence on the present. Additionally, if the internal energy of the material is affected by past random effects, it can be modeled as fractionally integrated additive noise, represented as 0Iγt˙W(t) using the classical Wiener process. Here, 0Iγtu (or RLD−γ0,tu) denotes the Riemann-Liouville fractional integral of order γ of the function u defined by
0Iγtu(t)≡RLD−γ0,tu(t)=1Γ(γ)∫t0(t−σ)γ−1u(σ)dσ. |
In this paper, we consider the time fractional semilinear stochastic partial differential equation driven by fractionally integrated additive noise,
{CDα0,tu(t)+Au(t)=f(u(t))+0Iγt˙W(t),0<t≤T,u(0)=u0, | (1.1) |
where 0<α<1, 0≤γ≤1, A=−Δ is a self-adjoint, positive definite, not necessarily bounded operator on the Hilbert space with domain D(A)=H2(D)∩H10(D), where D⊂Rd,d=1,2,3 denotes a bounded convex polygonal domain. Here, ˙W(t)=dW(t)dt denotes the white noise, and the time fractional derivative CDα0,tu with order α∈(0,1) is defined as follows [12],
CDα0,tu(t)=1Γ(1−α)∫t0(t−σ)−α∂u∂σdσ, |
where Γ(⋅) denotes the Gamma function.
In recent years, the numerical solution of fractional partial differential equations has become a major focus of research. This is because analytical expressions for such equations are generally difficult to obtain, prompting researchers to explore numerical methods instead. The presence of singular convolution kernels in fractional operators further complicates the solving process. To tackle this challenge, a plethora of remarkable mathematical techniques and innovative approaches have emerged in recent years. These techniques not only play a crucial role in theoretical investigations but also demonstrate great potential in practical applications. For example, a novel technique employing double reduction order and a newly constructed nonlinear compact difference operator has been developed to simulate nonlocal problems on graded meshes [13]. In a separate study, researchers [15] have devised a conservative, positivity-preserving, nonlinear finite volume scheme suitable for multi-term nonlocal Nagumo-type equations using distorted meshes. Additionally, [16] proposed a positivity-preserving finite volume scheme tailored for subdiffusion equations on nonconforming quadrilateral distorted meshes with hanging nodes. Furthermore, [20] addresses the numerical solution of the three-dimensional nonlocal evolution equation with a weakly singular kernel.
Many researchers are also dedicated to studying techniques for solving stochastic partial differential equations. Interested readers are encouraged to explore [5,6,14,17] for further works in this area. In prior research, our focus revolved around weak convergence analysis of the L1 scheme for a stochastic subdiffusion problem, as well as leveraging the spectral method for strong approximation of stochastic semilinear subdiffusion and superdiffusion equations driven by fractionally integrated additive noise [8,9]. In this article, we delve deeper into the analysis of non-smooth data in this model. Notably, our model problem (1.1), in contrast to the formulation studied in [6,14], exhibits greater generality and necessitates the involvement of three distinct Mittag-Leffler solution operators (namely E(t), ¯E(t), and ˜E(t), as detailed in Section 2) due to the presence of time fractional derivative and fractionally integrated additive noise.
Upon applying the Riemann-Liouville derivative operator RLD1−α0,t:=(0Iαt)′ on both sides of (1.1), it is then formally equivalent to a semilinear fractional Volterra type evolution equation
du(t)+RLD1−α0,tAu(t)dt=RLD1−α0,tf(u(t))dt+RLD1−α−γ0,tdW(t), | (1.2) |
so the existence and uniqueness of a mild solution u can be proved according to the literature methods [1] analogously, even only under some assumptions, via a standard Banach fixed point argument.
The main contributions of the paper are the following:
(i) The paper introduces finite element analysis for semilinear stochastic subdiffusion problems driven by fractionally integrated additive noise. It explores the smoothness of the solution and employs complex integration techniques to approximate the error of the solution operator under non-smooth data.
(ii) The paper establishes the continuity conditions of the mild solution, providing insights into the behavior and regularity of the solution when dealing with non-smooth data. We accurately demonstrate the impact of the fractional parameters α and γ on the convergence rates through numerical examples, offering valuable insights into the sensitivity and dependence of the solution on these parameters.
The remaining sections of this paper are structured as follows. In the upcoming section, we lay out the framework for the stochastic partial differential equation (SPDE) and establish significant smoothing properties using Mittag-Leffler functions. Subsequently, we derive essential error estimates for deterministic subdiffusion under non-smooth data, followed by obtaining optimal convergence estimates for the finite element method in the presence of non-smooth data.
It is known that if A=−Δ with homogenous Dirichlet boundary conditions, one has Aφk=λkφk,k∈N, where 0<λ1≤λ2≤⋯≤⋯≤λk≤⋯, limk→∞λk=∞, and the eigenvectors {φk}∞k=1 form an orthonormal basis for H. Let H=L2(D) be a separable Hilbert space with inner product (⋅,⋅) and norm ‖⋅‖. Let (Ω,F,P,{Ft}t≥0) be a filtered probability space, with Bochner spaces Lp(Ω;H)=Lp((Ω,F,P);H). Let E denote the expectation (with respect to P). We recall an abstract framework to describe the noise W(t) in the model (1.1) more precisely. A Wiener process W(t) with a covariance operator Q may be characterized by the Fourier type series as follows:
W(t)=∞∑k=1μ12kφkβk(t), | (2.1) |
where Q is a bounded, linear, self-adjoint, positive definite operator on H, with the pairs of eigenvalue and eigenfunction {(μk,φk)}∞k=1. The {βk(t)}∞k=1 is a sequence of independently and identically distributed standard Brownian motions.
For any ν∈R, we introduce the space ˙Hν(D)=D(Aν2) with norm |v|2ν=‖Aν2v‖2=∞∑k=1λνk(v,φk)2, where {φk}∞k=1 is the orthonormal basis in H.
Let L=L(H) denote the space of all bounded linear operators on H and let L02=HS(Q12(H),H) be the space of Hilbert-Schmidt operators from Q12(H) to H, i.e.,
L02={T∈L(H):∞∑k=1‖TQ12φk‖2<∞}, |
furnished with the norm ‖T‖2L02=∞∑k=1‖TQ12φk‖2, thus ‖T‖L02=‖TQ12‖HS<∞, for T∈L02.
We also need to recall the Burkholder-Davis-Gundy inequality [11], for p≥2,
‖∫t0ϕ(σ)dW(σ)‖Lp(Ω;H)≤Cp‖(∫t0‖ϕ(σ)‖2L02dσ)12‖Lp(Ω;R), | (2.2) |
for strongly measurable functions ϕ:[0,T]→L02. It's important to emphasize that when p=2, it is the Ito isometry.
By using time fractional Duhamel's principle and Laplace transform, we can obtain the mild solution of (1.1) as follows:
u(t)=E(t)u0+∫t0¯E(t−σ)f(u(σ))dσ+∫t0˜E(t−σ)dW(σ),P−a.s., | (2.3) |
where
E(t):=Eα,1(−tαA), | (2.4) |
¯E(t):=tα−1Eα,α(−tαA), | (2.5) |
˜E(t):=tα+γ−1Eα,α+γ(−tαA). | (2.6) |
The two parameter function of the Mittag-Leffler type plays a very important role in the fractional calculus [5]. We recall the following important properties of the Mittag-Leffler function essential in our analysis.
Lemma 2.1. Let 0<α<2 and β∈R be arbitrary and πα2<μ<min(π,απ), then there exists a constant C=C(α,β,μ) such that
|Eα,β(z)|≤{C(1+|z|)−1,β−α∉Z−,C(1+|z|)−2,β−α∈Z−,μ≤|arg(z)|≤π, | (2.7) |
where the notation Z− denotes the set of nonpositive integers, i.e., Z−={0,−1,−2,…}.
Throughout we always make the following standing assumption on the fractional orders α and γ, which is sufficient to ensure the well-posedness of Eq (1.1) (see [2,5]).
Assumption 2.1. Let 0<α<1, 0≤γ≤1 with α+γ>12.
On the Assumption 2.1, for β∈(0,κ], the regularity of noise can be characterized as follows (see Lemma A.1 of [5]):
‖Aβ−κ2‖L02=‖Aβ−κ2Q12‖HS≤C, | (2.8) |
where (with ε>0 small)
κ={2,if12<γ<1,2−ε,ifγ=12,2−1−2γα−ε,if0≤γ<12. | (2.9) |
Now we state the smoothing property of the operators E(t), ¯E(t), and ˜E(t).
Lemma 2.2. There exists C such that for t>0, we have
‖AsE(t)‖≤Ct−αs,s∈[0,1], | (2.10) |
‖A−s˙E(t)‖≤Ctαs−1,s∈[0,1], | (2.11) |
‖As¯E(t)‖≤Ct(1−s)α−1,s∈[0,1], | (2.12) |
‖A−s˙¯E(t)‖≤Ctα−2,s≥0, | (2.13) |
‖As˜E(t)‖≤Ct(1−s)α+γ−1,s∈[0,1], | (2.14) |
‖A−s˙˜E(t)‖≤Ctα+γ−2,s≥0. | (2.15) |
Proof. We just prove (2.13). The other conclusions can be obtained by using the similar method. By Lemma 2.1, we have
‖A−s˙¯E(t)‖=‖A−stα−2Eα,α−1(−Atα)‖≤supλ>0,t≥0λ−stα−2(1+tαλ)2≤Ctα−2. |
This completes the proof of the lemma.
Remark 2.1. Overall, the corresponding conclusions are the smoothing properties of the heat semigroup, when α→1 and γ→0, see [17].
In the qualitative theory of nonlinear PDEs, the subsequent Gronwall type inequalities play a very important role in error estimate.
Lemma 2.3. [3] Let T>0, N∈N, k=TN, and tn=nk for 0≤n≤N. If ζ1,...,ζN≥0 satisfy for some M0,M1≥0, and μ,ν>0, the inequality
ζn≤M0(1+t−1+μn)+M1kn−1∑j=1t−1+νn−jζj,1≤n≤N, |
then there exists a constant M2=M2(μ,ν,M1,T) such that ζn≤M0M2(1+t−1+μn),1≤n≤N.
To mimic Assumption 2.14 as stated in book [7], let PT be the σ-field of predictable stochastic processes, and B(S) be the Borel σ-field of S. Given the available options, we shall make some reasonable assumptions about the nonlinear parts.
Assumption 2.2. The mapping f:[0,T]×Ω×H→˙H−1, (t,ω,h)→f(t,ω,h) is PT×B(H)/B(˙H−1) measurable. When δ∈(0,12), there exists a constant C that satisfies the following expression:
‖f(t1,ω,h)−f(t2,ω,h)‖−1≤C(1+‖h‖)(t2−t1)δ, | (2.16) |
for all h∈H, 0≤t1<t2≤T, ω∈Ω.
In this section, we formulate the Galerkin finite element methods for spatial discretization in combination with time discretization based on an exponential Euler type method for approximation of (1.1). Let Th be a regular shaped quasi-uniform triangulation of the domain D, and let Sh⊂H10(D) be the space of continuous piecewise linear functions on the triangulation Th. We define the L2-projection Ph:H→Sh by
(Phu,χ)=(u,χ),χ∈Sh, | (3.1) |
and the Ritz projection Rh:H10→Sh by
a(Rhu,χ)=a(u,χ),χ∈Sh, |
where a(u,χ)=(∇u,∇χ) is the associated bilinear form.
Note that by interpreting the righthand side of (3.1) as a duality pairing between ˙H1(D) and ˙H−1(D), one may extend Ph to be a bounded operator from ˙H−1(D) to Sh. It is well-known that the operators Ph and Rh have the following approximation properties [18].
Lemma 3.1. The operators Ph and Rh satisfy
‖Phu−u‖+h‖∇(Phu−u)‖≤Chq|u|q,foru∈˙Hq,q=1,2,‖Rhu−u‖+h‖∇(Rhu−u)‖≤Chq|u|q,foru∈˙Hq,q=1,2. |
The semi-discrete Galerkin FEM scheme for (1.1) is to find uh(t)∈Sh such that
{CDα0,tuh(t)+Ahuh(t)=Phf(uh(t))+0IγtPh˙W(t),0<t≤T,uh(0)=Phu0, | (3.2) |
where the discrete Laplacian Ah is defined by
Ah:Sh→Sh,(Ahψ,χ)=a(ψ,χ),∀ψ,χ∈Sh. |
Naturally, we present the discrete analogues of operators Eh(t), ¯Eh(t), and ˜Eh(t) as follows
Eh(t):=Eα,1(−tαAh), | (3.3) |
¯Eh(t):=tα−1Eα,α(−tαAh), | (3.4) |
˜Eh(t):=tα+γ−1Eα,α+γ(−tαAh). | (3.5) |
Now, let 0<t0<t1<⋯<tM=T be a uniform partion of time interval [0,T], with time step Δt=tm+1−tm,m=0,1,⋯,M−1. Hence, the fully discrete approximation of (1.1) is given by
Umh=Eh(tm)Phu0+m−1∑j=0∫tj+1tj¯Eh(tm−σ)dσ(PhF(Ujh))+∫tm0˜Eh(tm−σ)PhdW(σ), | (3.6) |
with initial value U0h=Phu0.
Let us introduce and prove some Lemmas that will play an important role later on.
Lemma 3.2. [10] Let 0≤ω≤μ≤2. For α∈(0,1), there exists a constant C such that
‖(E(t)−Eh(t)Ph)v‖≤Chμt−αμ−ω2‖v‖ω,forv∈˙Hω. |
We now turn to the non-smooth data error estimates of the approximations to ¯E(t)g,g∈H in the semi-discrete case.
Lemma 3.3. Let 0≤s≤1 and 0≤r≤2 with r+s≤2. For g∈H, there holds
‖As2(¯E(t)−¯Eh(t)Ph)g‖≤Ch2−s−rtαr2−1‖g‖. | (3.7) |
Proof. In the case s=0, by the inverse Laplace transform, for any given g∈H, we have
¯E(t)g=12πi∫Γθ,δezt(zα+A)−1gdz, | (3.8) |
¯Eh(t)Phg=12πi∫Γθ,δezt(zα+Ah)−1Phgdz | (3.9) |
where Γτθ,δ={z∈Γθ,δ:|ℑz|≤πτ}, and Γθ,δ={z∈C:z=re±iθ,r≥δ}∪{z∈C:z=δeiϕ,|ϕ|≤θ},π2<θ<π, πτ>δ.
Let us first show (3.7). For any fixed g∈H, we have, by (3.8) and (3.9),
‖(¯E(t)−¯Eh(t)Ph)g‖≤C∫Γθ,δeℜ(z)t‖((zα+A)−1−(zα+Ah)−1Ph)g‖|dz|. |
By [4], p. 820, we get ‖((zα+A)−1−(zα+Ah)−1Ph)g‖≤Ch2‖g‖,∀z∈Γθ,δ. Hence, we have
‖(¯E(t)−¯Eh(t)Ph)g‖≤Ch2‖g‖(∫{z∈C:z=δeiϕ,|ϕ|≤θ}+∫{z∈C:z=re±iθ,r≥δ})eℜ(z)t|dz|=I+II. |
For I, with z=δeiϕ, we have δ=t−1, where t−1<πτ for sufficiently small τ,
I≤Ch2‖g‖∫θ−θetδcosϕδdϕ≤Ch2‖g‖δ∫θ−θecosϕdϕ≤Ch2t−1‖g‖. |
For II, with z=re±iθ, we have r≥δ,δ=t−1,
II≤Ch2‖g‖∫∞δetrcosθdr≤Ch2‖g‖∫∞t−1e−ctrdr≤Ch2‖g‖t−1∫∞ce−xdx≤Ch2t−1‖g‖, |
where c,C is some suitable positive constant.
Meanwhile, by (2.12) and the triangle inequality,
‖(¯E(t)−¯Eh(t)Ph)g‖≤Ctα−1‖g‖. |
Similarly, for s=1, there holds
‖A12(¯E(t)−¯Eh(t)Ph)g‖≤Cht−1‖g‖. |
Now the desired assertion follows by interpolation, which completes the proof.
The next lemma gives an error estimate on ˜E(t). More details can be found in Lemma 4.4 of [5].
Lemma 3.4. Let 0≤s≤1 and 0≤r≤2 with r+s≤2. For g∈H, there holds
‖As2(˜E(t)−˜Eh(t)Ph)g‖≤Ch2−s−rtαr2+γ−1‖g‖. | (3.10) |
Based on the previous discussion, we are ready to prove the error estimates for the fully discrete approximation.
Theorem 3.1. For 0≤r≤2 and rα+2γ>1, by the Assumptions 2.1 and 2.2, with ν∈[0,β), β∈(0,κ], α+γ∈(12,1) and α(2−κ+β−ν)+2γ−1∈(0,1), then there holds
suptm∈[0,T]‖u(tm)−Umh‖L2(Ω;H)≤C(h(2−r)α+h2−rmax{t−(2−r)α2m,lntmh2−r}+Δtαν). |
Proof. Due to (1.1), it is formally then equivalent to a nonlinear fractional Volterra type evolution equation. We can obtain supt∈[0,T]‖u(t)‖Lp(Ω;˙Hν)≤C. Subtracting (3.6) from (2.3) and by taking norms, one obtains
‖u(tm)−Umh‖L2(Ω;H)≤‖E(tm)u0−Eh(tm)Phu0‖L2(Ω;H)+‖∫tm0(ˉE(tm−σ)−ˉEh(tm−σ)Ph)F(u(σ))dσ‖L2(Ω;H)+‖∫tm0(˜E(tm−σ)−˜Eh(tm−σ)Ph)dW(σ)‖L2(Ω;H)+‖m−1∑j=0∫tj+1tjˉE(tm−σ)Ph(F(u(σ))−F(Ujh))dσ‖L2(Ω;H)=I1+I2+I3+I4. |
Here, we note that I1, I2, and I3 correspond to the spatial finite element discretization error, while I4 corresponds to the temporal error.
The estimate of I1 is a consequence of Lemma 3.2. For ω=0 and μ=2−r, we get
I1=‖E(tm)u0−Eh(tm)Phu0‖L2(Ω;H)≤Ch2−rt−(2−r)α2m. |
For I2, by using (3.7) and r>0, we have
I2=‖∫tm0(¯E(tm−σ)−¯Eh(tm−σ))PhF(u(σ))dσ‖L2(Ω;H)≤∫tm0‖¯E(tm−σ)−¯Eh(tm−σ)‖⋅(1+‖u(σ)‖L2(Ω;H))dσ≤C∫tm0‖¯E(tm−σ)−¯Eh(tm−σ)‖dσ=C∫tm0‖¯E(σ)−¯Eh(σ)‖dσ≤∫tm0Ch2−rσrα2−1dσ=Ch2−rtrα2m. |
In the case r=0, similarly, we split I2 into two terms,
I2≤∫h2−r0‖¯E(σ)−¯Eh(σ)‖dσ+∫tmh2−r‖¯E(σ)−¯Eh(σ)‖dσ=I21+I22. |
For I21, noting ‖¯E(σ)‖≤Cσα−1 and ‖¯Eh(σ)‖≤Cσα−1, then
I21≤∫h2−r0‖¯E(σ)−¯Eh(σ)‖dσ≤Ch(2−r)α. |
For I22, by (3.7) with r=0, we derive
I22≤∫tmh2−r‖¯E(σ)−¯Eh(σ)‖dσ≤∫tmh2−rCh2−rσ−1dσ=Ch2−rlh, |
where lh=lntmh2−r.
Now, we estimate I3. Using (3.10), by Itô's formula, for trace class and rα+2γ>1, we obtain
I23=‖∫tm0(˜E(tm−σ)−˜Eh(tm−σ)Ph)dW(σ)‖2L2(Ω;H)=∫tm0‖(˜E(tm−σ)−˜Eh(tm−σ)Ph)‖2L02dσ≤C∫tm0‖(˜E(σ)−˜Eh(σ))‖2L02dσ≤C∫tm0‖(˜E(σ)−˜Eh(σ))Q12‖2HSdσ≤C∫tm0‖˜E(σ)−˜Eh(σ)‖2⋅‖Q12‖2HSdσ≤C∫tm0(h2−rσαr2+γ−1)2⋅‖Q12‖2HSdσ≤C∫tm0(h2−rσrα2+γ−1)2dσ≤Ch4−2rtrα+2γ−1m. |
For I4, by Lemma 2.2, the idea is similar to [9] with different coefficient condition α(2−κ+β−ν)+2γ−1>0, so we get I4≤Δtαν. Finally, combining the Gronwall's inequality of Lemma 2.3, we complete the proof of Theorem 3.1.
Under the assumptions of smoothness of the solution operator and the nonlinear term, we can obtain the following conclusions.
Theorem 3.2. Let p∈[2,∞) be given such that Assumptions 2.13–2.17 hold, then the unique mild solution u to (2.3) is continuous with respect to the norm ‖⋅‖Lp(Ω;˙Hs).
Proof. According to the definition of continuity, our goal is to show that
limt2−t1→0t1<t2‖u(t2)−u(t1)‖Lp(Ω;˙Hs)=0 |
with either t1 or t2 fixed.
Therefore, with the expression of a mild solution u, we divide ‖u(t2)−u(t1)‖Lp(Ω;˙Hs) into five parts, and only need to show that each of the parts approaches zero when t2→t1. By utilizing the triangle inequality, we obtain
‖u(t1)−u(t2)‖Lp(Ω;˙Hs)≤‖E(t1)u0−E(t2)u0‖Lp(Ω;˙Hs)+‖∫t2t1ˉE(t2−σ)f(u(σ))dσ‖Lp(Ω;˙Hs)+‖∫t10(ˉE(t2−σ)−ˉE(t1−σ))f(u(σ))dσ‖Lp(Ω;˙Hs)+‖∫t2t1˜E(t2−σ)dW(σ)‖Lp(Ω;˙Hs)+‖∫t10(˜E(t2−σ)−˜E(t1−σ))dW(σ)‖Lp(Ω;˙Hs)=M1+M2+M3+M4+M5. |
For the expression M1, we apply (2.11) from Lemma 2.2. One yields
M1=‖E(t1)u0−E(t2)u0‖Lp(Ω;˙Hs)=‖As2(E(t1)−E(t2))u0‖Lp(Ω;H)=‖∫t2t1As2˙E(τ)u0 dτ‖Lp(Ω;H)=‖∫t2t1A−v2˙E(τ)Av+s2u0 dτ‖Lp(Ω;H)≤‖u0‖Lp(Ω;˙Hs+v)⋅∫t2t1‖A−v2˙E(τ)‖dτ≤C‖u0‖Lp(Ω;˙Hs+v)⋅∫t2t1ταv2−1 dτ≤C‖u0‖Lp(Ω;˙Hs+v)⋅(tαv22−tαv21), |
and the validity of limt2−t1M1=0 here is obvious when ν≥0.
For M2, we insert a node t3 and split it into three parts.
M2=‖∫t2t1¯E(t2−σ)f(u(σ))dσ‖Lp(Ω;˙Hs)≤‖∫t2t1¯E(t2−σ)f(u(σ))dσ‖Lp(Ω;˙Hs)≤‖∫t2t1As+12¯E(t2−σ)A−12(f(u(σ))−f(u(t2)))dσ‖Lp(Ω;H)+‖∫t2t1As+12¯E(t2−σ)A−12(f(u(t2))−f(u(t3)))dσ‖Lp(Ω;H)+‖∫t2t1As+12¯E(t2−σ)A−12f(u(t3))dσ‖Lp(Ω;H)=M21+M22+M23. |
For M21, we apply (2.12) from Lemma 2.2, and Assumption 2.2. We obtain
M21≤∫t2t1‖As+12¯E(t2−σ)A−12(f(u(t2))−f(u(σ)))‖Lp(Ω;H)dσ≤C∫t2t1(t2−σ)−s+12α+α−1(t2−σ)δ(1+‖u(σ)‖Lp(Ω;H))dσ≤C(t2−t1)1−s2α+δ(1+supσ∈[0,T]‖u(σ)‖Lp(Ω;H)). |
Due to δ∈(0,12), therefore 1−s2α+δ>0, and thus when t2→t1, it is valid.
For M22,
M22=‖∫t2t1As+12¯E(t2−σ)A−12(f(u(t2))−f(u(t3)))dσ‖Lp(Ω;H)≤∫t2t1‖As+12¯E(t2−σ)A−12(f(u(t2))−f(u(t3)))‖Lp(Ω;H)dσ, |
which is formally identical to M21. When t2→t3∈[t1,t2] holds, the conclusion is valid.
For M23, with the help of (2.12), one gets
M23=‖∫t2t1As+12¯E(t2−σ)A−12f(u(t3))dσ‖Lp(Ω;H)≤∫t2t1‖As+1−r2¯E(t2−σ)A−1+r2f(u(t3))‖Lp(Ω;H)dσ≤C∫t2t1(t2−σ)1+r−s2αdσ⋅(1+supσ∈[0,T]‖Y(σ)‖Lp(Ω;˙Hr)). |
To ensure the continuity property holds, it is sufficient to only have the integration in the last inequality be well-defined and satisfy the conditions of 1+γ−s2α>0.
For M3, we prove that the following property holds first. One obtains
‖As˙¯E(t)‖2=||Astα−2Eα,α−1(−Atα)||2=|tα−2Eα,α−1(−Atα)|2s=∞∑j=1λ2sjt2(α−2)⋅Eα,α−1(−tαλj)2⋅(v,φj)2≤t2[(1−s)α−2]⋅∞∑j=1(tαλj)2s(1+t4λj)4(v,φj)2, |
and then we have
M3=‖∫t10(¯E(t2−σ)−¯E(t1−σ))f(u(σ))dσ‖Lp(Ω;˙Hs)≤∫t10‖(¯E(t2−σ)−¯E(t1−σ))f(u(σ))‖Lp(Ω;˙Hs)dσ=∫t10‖∫t2t1As+1−r2˙¯E(t2−σ)dτA−1+r2f(u(σ))‖Lp(Ω;H)dσ≤C∫t2t1|(t2−σ)1+r−s2α−1−(t1−σ)1+r−s2α−1|dσ⋅(1+supσ∈[0,T]‖u(σ)‖Lp(Ω;˙Hr))≤C(t2−t1)1+r−s2α⋅(1+supσ∈[0,T]‖u(σ)‖Lp(Ω;˙Hr)). |
Since 1+r−s2α>0 holds here, the conclusion is valid.
For M4, by applying (2.14) and the Ito formula, we can obtain
M4=‖∫t2t1˜E(t2−σ)dW(σ)‖Lp(Ω;˙Hs)≤C‖(∫t2t1‖As2˜E(t2−σ)‖2L02dσ)12‖Lp(Ω;R)≤C‖(∫t2t1(t2−σ)2(−s2α+α+γ−1)dσ)12‖Lp(Ω;R)≤C(t2−t1)(1−s2)α+γ−12. |
In order for the conclusion to hold, it is necessary for the conditions of (1−s2)α+γ−12>0 to be satisfied here.
For M5, according to the requirement, we first prove that the following property holds.
‖As˙˜E(t)‖2≤‖Astα+γ−2Eα,α+γ−1(−Atα)‖2=|tα+γ−2Eα,α+γ−1(−Atα)|22s=∞∑j=1λ2sjt2(α+γ−2)Eα,α+γ−1(−tαλj)2(v,φ)2≤t2[(1−s)α+γ−2]∞∑j=1(λjtα)2s(1+tαλj)2(v,φ)2, |
then we have
M5=‖∫t10(˜E(t2−σ)−˜E(t1−σ))dW(σ)‖Lp(Ω;˙Hs)≤C‖(∫t10||As2(˜E(t2−σ)−˜E(t1−σ))‖2L02dσ)12‖Lp(Ω;R)=C‖(∫t10∫t2t1‖As2˙˜E(τ−σ)‖2L02dσ)12‖Lp(Ω;R)≤C‖(∫t10∫t2t1[(τ−σ)(1−s2)α+γ−2]2dτdσ)12‖Lp(Ω;R)=C‖(∫t10[(t2−σ)(2−s)α+2γ−3−(t1−σ)(2−s)α+2γ−3]dσ)12‖Lp(Ω;R)≤C(t2−t1)(1−s2)α+γ−1. |
The last term of the inequality only needs to satisfy the condition of (1−s2)α+γ−12>0. Thus, the theorem is proved.
We conducted several numerical experiments in this section to validate our previous theoretical findings. To illustrate the theoretical results obtained in Theorem 3.1 for α∈(0,1), we provide two numerical examples. Specifically, we set T=0.1, D=(0,1), and F(u)=sin(u).
To explain the computer implementation of the full-discrete method (3.6), we make the assumption that the covariance operator Q possesses the same eigenfunctions as A, such that Qv=∞∑k=1μk(v,φk)φk. Furthermore, we assume that W(t) has the following Fourier series expansion: W(t)=∑∞k=1μ1/2kφkβk(t). Thus, the semi-discrete solution UNm,k,h satisfies
UNm,k,h=Eα,1(−tαmAh)Phu0,k+m−1∑j=0∫tj+1tj(tm−σ)α−1Eα,α(−(tm−σ)αAh)dσ⋅PhFk(u(tj))+∫tm0(tm−σ)α+γ−1Eα,α+γ(−(tm−σ)αAh)Phμ12kdβk(σ), |
where u0,k=(u0,φk), Fk(⋅)=(F(⋅),φk), and βk(σ) for k=1,…,N are mutually independent standard Brownian motions.
In our experiments, we investigated the dependence of the error estimates in Theorem 3.1 on the time step size Δt. To approximate the exact solution, we used the fully discrete solution with a small time step size of Δt=2−10 and a spatial step size of Δx=2−10. The theoretical rate of convergence is αν for ν∈(0,1), which approaches α as ν→1.
To conduct the experiments, we fixed a small spatial step size Δx and considered a sequence of moderate time step sizes Δti=2−i,i=2,…,5. We performed M=100 simulations for each time step size Δti. In each simulation ωj, j=1,2,…,M, we generated Nh independent Brownian motions βk(t), k=1,2,…,Nh. The initial data was set to u0=1, and we defined the nonlinear operator F(u)=sin(u).
In Figure 1, we investigated the effects of different parameters α and γ. We observed that the numerical results are consistent with the theoretical results stated in Theorem 3.1. The numerical values vary slightly due to the limited range of α and γ. As the mesh size is refined, the numerical experiments support the theoretical findings in Theorem 3.1.
In the following analysis, we focus on the spatial convergence. To this end, we consider a fixed number of space steps M=200 and final time T=1, and we obtain the reference solution at N=480. Next, we compute the numerical results for different combinations of fractional orders α and γ, with trace class noise (with m=2). Table 1 illustrates the results, which show that an O(h2) convergence order is observed for all combinations. We observed that the initial convergence rate changes slowly, but as the mesh is refined, the results are in excellent agreement with the theoretical predictions from Theorem 3.1.
γ | α∖M | 10 | 20 | 40 | 80 | 160 | order |
0.2 | 0.3 | 7.6512e-03 | 2.0123e-03 | 5.1534e-04 | 1.2745e-04 | 2.9646e-05 | 2.00 (2.00) |
0.7 | 4.8203e-03 | 1.3214e-03 | 3.4625e-04 | 8.7435e-05 | 2.0347e-05 | 1.97 (2.00) | |
0.6 | 0.3 | 2.3923e-03 | 6.2545e-04 | 1.5964e-04 | 3.9323e-05 | 9.0951e-06 | 2.01 (2.00) |
0.7 | 2.2653e-03 | 5.9214e-04 | 1.5031e-04 | 3.7334e-05 | 8.6453e-06 | 2.00 (2.00) | |
0.8 | 0.3 | 2.0211e-03 | 5.2732e-04 | 1.3312e-04 | 3.3132e-05 | 7.6553e-06 | 2.01 (2.00) |
0.7 | 2.0132e-03 | 5.2513e-04 | 1.3321e-04 | 3.3051e-05 | 7.6249e-06 | 2.01 (2.00) |
In this study, we proposed a numerical method for solving stochastic semilinear subdiffusion equations driven by fractionally integrated additive noise. The temporal discretization relies on Mittag-Leffler integrators, while the spatial discretization is based on the finite element method. The effectiveness of the proposed method was demonstrated through illustrative examples that provided support for the theoretical analysis. In the next phase, we aim to enhance efficiency by incorporating the stochastic fractional system model. We also plan to preserve important physical properties and structures, such as positivity preservation, maximum principle, long time behavior, and investigate singular solutions. This includes studying the uniform L1 long time behavior of time discretization for time-fractional partial differential equations with non-smooth data. Additionally, we will develop a finite volume scheme for two-dimensional time-fractional stochastic Fokker-Planck equations on distorted meshes, which preserves the maximum principle.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work is supported by the NSFC under Grant No. 12147115, the University Natural Science Research Project of Anhui Province under Grant No. 2023AH050254, the Fundamental Research Program of Shanxi Province under Grant No. 202103021224317, the Lvliang High Tech Research and Development Program under Grant No. 2023GXYF14, and the present job was also partially supported by the School Research Project of AUFE, Grant No. ACKYC21049.
The authors declare no conflict of interest.
[1] |
B. Baeumer, M. Geissert, M. Kovács, Existence, uniqueness and regularity for a class of semilinear stochastic Volterra equations with multiplicative noise, J. Differ. Equations, 258 (2015), 535–554. https://doi.org/10.1016/j.jde.2014.09.020 doi: 10.1016/j.jde.2014.09.020
![]() |
[2] |
Z. Q. Chen, K. H. Kim, P. Kim, Fractional time stochastic partial differential equations, Stoch. Proc. Appl., 125 (2015), 1470–1499. https://doi.org/10.1016/j.spa.2014.11.005 doi: 10.1016/j.spa.2014.11.005
![]() |
[3] | C. M. Elliott, S. Larsson, Error estimates with smooth and nonsmooth data for a finite element method for the Cahn-Hilliard equation, Math. Comp., 58 (1992), 603–630. |
[4] | H. Fujita, T. Suzuki, Evolution problems, In Handbook of Numerical Analysis, North-Holland, Amsterdam, 2 (1991), 789–928. |
[5] |
B. T. Jin, Y. B. Yan, Z. Zhou, Numerical approximation of stochastic time-fractional diffusion, ESAIM: M2AN, 53 (2019), 1245–1268. https://doi.org/10.1051/m2an/2019025 doi: 10.1051/m2an/2019025
![]() |
[6] |
M. Kovács, S. Larsson, F. Saedpanah, Mittag-Leffler Euler integrator for a stochastic fractional order equation with additive noise, SIAM J. Numer. Anal., 58 (2020), 66–85. https://doi.org/10.1137/18M1177895 doi: 10.1137/18M1177895
![]() |
[7] | R. Kruse, Strong and weak approximation of semilinear stochastic evolution equations, Springer, Berlin, 2016. |
[8] |
Y. Hu, C. P. Li, Y. Yan, Weak convergence of the L1 scheme for a stochastic subdiffusion problem driven by fractionally integrated additive noise, Appl. Numer. Math., 178 (2022), 192–215. https://doi.org/10.1016/j.apnum.2022.04.004 doi: 10.1016/j.apnum.2022.04.004
![]() |
[9] | Y. Hu, Y. Yan, Shahzad Sarwar, Strong approximation of stochastic semilinear subdiffusion and superdiffusion driven by fractionally integrated additive noise, Numer. Meth. Part. D. E., 2023. https://doi.org/10.1002/num.23068 |
[10] |
X. C. Li, X. Y. Yang, Error estimates of finite element methods for stochastic fractional differential equations, J. Comput. Math., 35 (2017), 346–362. https://doi.org/10.4208/jcm.1607-m2015-0329 doi: 10.4208/jcm.1607-m2015-0329
![]() |
[11] | G. D. Prato, J. Zabczyk, Stochastic equations in infinite dimensions, Cambridge University Press, Cambridge, 1992. https://doi.org/10.1017/CBO9780511666223 |
[12] | I. Podlubny, Fractional differential equations, Academic Press, San Diego, 1999. |
[13] |
J. W. Wang, X. X. Jiang, X. H. Yang, H. X. Zhang, A nonlinear compact method based on double reduction order scheme for the nonlocal fourth-order PDEs with Burgers' type nonlinearity, J. Appl. Math. Comput., 70 (2024), 489–511. https://doi.org/10.1007/s12190-023-01975-4 doi: 10.1007/s12190-023-01975-4
![]() |
[14] |
X. J. Wang, R. S. Qi, A note on an accelerated exponential Euler method for parabolic SPDEs with additive noise, Appl. Math. Lett., 46 (2015), 31–37. https://doi.org/10.1016/j.aml.2015.02.001 doi: 10.1016/j.aml.2015.02.001
![]() |
[15] |
X. H. Yang, Z. M. Zhang, On conservative, positivity preserving, nonlinear FV scheme on distorted meshes for the multi-term nonlocal Nagumo-type equations, Appl. Math. Lett., 150 (2024), 108972. https://doi.org/10.1016/j.aml.2023.108972 doi: 10.1016/j.aml.2023.108972
![]() |
[16] |
X. H. Yang, Z. M. Zhang, Q. Zhang, G. W. Yuan, Simple positivity-preserving nonlinear finite volume scheme for subdiffusion equations on general non-conforming distorted meshes, Nonlinear Dyn., 108 (2022), 3859–3886. https://doi.org/10.1007/s11071-022-07399-2 doi: 10.1007/s11071-022-07399-2
![]() |
[17] |
Y. Yan, Galerkin finite element methods for stochastic parabolic partial differential equations, SIAM J. Numer. Anal., 43 (2005), 1363–1384. https://doi.org/10.1137/040605278 doi: 10.1137/040605278
![]() |
[18] | V. Thomée, Galerkin finite element methods for parabolic problems, Springer-Verlag, Berlin, 2007. |
[19] | S. G. Samko, A. A. Kilbas, O. I. Marichev, Fractional integrals and derivatives: Theory and applications, Gordon and Breach, New York, 1993. |
[20] |
H. X. Zhang, Y. Liu, X. H. Yang, An efficient ADI difference scheme for the nonlocal evolution problem in three-dimensional space, J. Appl. Math. Comput., 69 (2023), 651–674. https://doi.org/10.1007/s12190-022-01760-9 doi: 10.1007/s12190-022-01760-9
![]() |
γ | α∖M | 10 | 20 | 40 | 80 | 160 | order |
0.2 | 0.3 | 7.6512e-03 | 2.0123e-03 | 5.1534e-04 | 1.2745e-04 | 2.9646e-05 | 2.00 (2.00) |
0.7 | 4.8203e-03 | 1.3214e-03 | 3.4625e-04 | 8.7435e-05 | 2.0347e-05 | 1.97 (2.00) | |
0.6 | 0.3 | 2.3923e-03 | 6.2545e-04 | 1.5964e-04 | 3.9323e-05 | 9.0951e-06 | 2.01 (2.00) |
0.7 | 2.2653e-03 | 5.9214e-04 | 1.5031e-04 | 3.7334e-05 | 8.6453e-06 | 2.00 (2.00) | |
0.8 | 0.3 | 2.0211e-03 | 5.2732e-04 | 1.3312e-04 | 3.3132e-05 | 7.6553e-06 | 2.01 (2.00) |
0.7 | 2.0132e-03 | 5.2513e-04 | 1.3321e-04 | 3.3051e-05 | 7.6249e-06 | 2.01 (2.00) |
γ | α∖M | 10 | 20 | 40 | 80 | 160 | order |
0.2 | 0.3 | 7.6512e-03 | 2.0123e-03 | 5.1534e-04 | 1.2745e-04 | 2.9646e-05 | 2.00 (2.00) |
0.7 | 4.8203e-03 | 1.3214e-03 | 3.4625e-04 | 8.7435e-05 | 2.0347e-05 | 1.97 (2.00) | |
0.6 | 0.3 | 2.3923e-03 | 6.2545e-04 | 1.5964e-04 | 3.9323e-05 | 9.0951e-06 | 2.01 (2.00) |
0.7 | 2.2653e-03 | 5.9214e-04 | 1.5031e-04 | 3.7334e-05 | 8.6453e-06 | 2.00 (2.00) | |
0.8 | 0.3 | 2.0211e-03 | 5.2732e-04 | 1.3312e-04 | 3.3132e-05 | 7.6553e-06 | 2.01 (2.00) |
0.7 | 2.0132e-03 | 5.2513e-04 | 1.3321e-04 | 3.3051e-05 | 7.6249e-06 | 2.01 (2.00) |