1.
Introduction
Fractional differential equations have wide applications in various fields of science, including physics, economics, engineering, chemistry, biology and others[1,2,3,4,5]. There are many kinds of definitions for the fractional derivatives, the most used fractional derivatives are the Riemann-Liouville fractional derivative and the Caputo fractional derivative [6,7]. However, both of these operators still present challenges in practical applications. To be more precise, the Riemann-Liouville derivative of a constant is non-zero and the Laplace transform of this derivative contains terms that lack physical significance. The Caputo fractional derivative has successfully addressed both issues, however, its definition involves a singular kernel which poses challenges in analysis and computation. Caputo and Fabrizio [8] have proposed a novel definition of the fractional derivative with a smooth kernel, referred to as the Caputo and Fabrizio (CF) derivatives, which present distinct representations for the temporal and spatial variables. The representation in time variable is suitable to use the Laplace transform, and the spatial representation is more convenient to use the Fourier transform. Although there is ongoing debate regarding the mathematical properties of fractional derivatives with non-singular kernels [9,10], numerous scholars remain interested in studying differential equations involving such derivatives due to their nice performance in various applications. Considering the CF derivative offers two primary advantages: 1) The utilization of a regular kernel in non-local systems is motivated by its potential to accurately depict material heterogeneities and fluctuations of various scales, which cannot be adequately captured by classical local theories or fractional models with singular kernels, see, e.g., [8,11]; 2) CF derivatives have numerical advantages. As we know, the truncation error of the numerical calculation for fractional operators with singular kernels is typically dependent on the order α. For instance, in the case of the Caputo fractional derivative, employing classical L1 discretization results in an error of order 2−α, which becomes highly unfavorable when α≈1. In order to enhance actuarial accuracy, the utilization of higher-order methods will lead to an increase in computational complexity, particularly for problems with high dimensions. However, within the same approximation framework, the CF derivative has a higher truncation error, see Remark 2.2. Further properties and diverse applications of this fractional derivative can be found in various references, such as [11,12,13,14,15,16,17].
Let T,L>0, Λ:=(0,S). In this paper, we are concerned with the numerical approximation of the multi-term time-fractional diffusion equation
with the initial conditions
and the boundary condition
where
0<α1≤α2≤⋯≤αn−1≤αn<1, and di≥0,i=1,2,⋯,n,n∈N. φ(x) and f(x,t) are given sufficiently smooth functions in their respective domains. In addition, CF0Dαitu(x,t) is the Caputo-Fabrizio derivative operator of order αi [8,11] defined as
If n=1, then (1.1)–(1.3) reduces to the single-term time-fractional diffusion equation. The model of (1.1)–(1.3), which describes the temporal flow of water within a leaky aquifer at various scales [12,18], as well as the electro-magneto-hydrodynamic flow of non-Newtonian biofluids with heat transfer [19], etc. For the well-posedness of (1.1)–(1.3), we refer to, e.g., [17,20,21].
Many researchers have explored the numerical approximation of both single-term and multi-term time fractional diffusion equations. In [22], Liu et al. proposed a finite difference method for solving time-fractional diffusion equations in both space and time domains. Lin and Xu [23] utilized a finite difference scheme in time and Legendre spectral methods in space to numerically solve the time-fractional diffusion equations. Subsequently, Li and Xu [24] improved upon their previous work by proposing a space-time spectral method for these equations. For the numerical treatment of multi-term time-fractional diffusion equations, [25] proposed a fully-discrete schemes for one- and two-dimensional multi-term time fractional sub-diffusion equations. These schemes combine the compact difference method for spatial discretization with L1 approximation for time discretization. The Galerkin finite element method and the spectral method were introduced in [26] and [27,28], respectively. Zhao et al. [29] developed a fully-discrete scheme for a class of two-dimensional multi-term time-fractional diffusion equations with Caputo fractional derivatives, utilizing the finite element method in spatial direction and classical L1 approximation in temporal direction. Akman et al. [30] proposed a numerical approximation called the L1-2 formula for the Caputo-Fabrizio derivative using quadratic interpolation. In [31], finite difference/spectral approximations for solving two-dimensional time CF fractional diffusion equation were proposed and analyzed. Later, a second order scheme [32] was devised for addressing this problem. A compact alternating direction implicit (ADI) difference scheme was proposed by [33] for solving the two-dimensional time-fractional diffusion equation.
Simulating models with fractional derivatives presents a challenge due to their non-locality, which significantly impedes algorithm efficiency and necessitates greater memory storage compared to traditional local models. In particular, for fractional models, the computational complexity of obtaining an approximate solution is O(N2T) and the required memory storage is O(NT), which contrasts with local models that have a complexity of O(NT) and require a memory storage of O(1), where NT denotes the total number of time steps, see, e.g., [23,31,32]. To address this issue, several researchers have proposed efficient algorithms for computing the derivatives of Riemann-Liouville, Caputo, and Riesz fractional operators, see e.g., [34,35,36,37] and the references therein. Recently, a fast compact finite difference method for quasi-linear time-fractional parabolic equations is presented and analyzed in [38]. Then, [39] proposed a fast second-order numerical scheme for approximating the Caputo-Fabrizio fractional derivative at node t=tk+1/2 with computational complexity of O(NT) and memory storage of O(1).
Inspired by the above mentioned, we extend finite difference/spectral approximations for the multi-term Caputo-Fabrizio time-fractional diffusion equation (1.1)–(1.3). Firstly, we present a L1 formula for the Caputo-Fabrizio derivative. In this context, we introduce two discrete fractional differential operators, namely Lαt and Fαt, which are essentially equivalent. However, Fαt effectively utilizes the exponential kernel and incurs lower storage and computational costs compared to Lαt. The idea of this approach is essentially identical to that of reference [39], albeit with a slightly different formulation in our case; specifically, the approximation is centered at point t=tk and presented in a more concise manner. The error bounds associated with these two operators will be examined in detail. Secondly, we develop a semi-discrete scheme based on finite difference method for multi-term time-fractional derivatives, with complete proofs of its unconditional stability and convergence rate. A detailed error analysis is carried out for the semi-discrete problem, showing that the temporal accuracy is second order. Finally, we present the fully-discrete scheme based on the Legendre spectral collocation method for spatial discretization. We will investigate both the convergence order of this method and its implementation efficiency, while providing a rigorous proof of its spectral convergence in this paper.
The rest of this paper is organized as follows. In Section 2, a semi-discrete scheme is proposed for (1.1)–(1.3) based on fast L1 finite difference scheme. The stability and convergence analysis of the semi-discrete scheme is presented. In Section 3, we construct a Legendre spectral collocation method for the spatial discretization of the semi-discrete scheme. Error estimates are provided for the full discrete problem. Some numerical results are reported in Section 4. Finally, the conclusions are given in Section 5.
2.
Semi-discretization
Define tk:=kΔt, k=0,1,⋯,NT, where Δt:=T/NT is the time step.
2.1. Fast L1 formula for Caputo-Fabrizio derivative
We first give L1 approximation for fractional Caputo-Fabrizio derivative of function h(t) defined by
In order to simplify the notations, we denote h(tk):=hk for 0≤k≤Mt. The L1 formula is obtained by substituting the linear Lagrange interpolation of h(t) into (2.1). Precisely, the linear approximation of the function h(t) on [tj−1,tj] is written as
and the error in the approximation is
Then we define the discrete fractional differential operator Lαt by
where bj,k:=σj,k−σj−1,k and
The right hand side of (2.4) involves a sum of all previous solutions {hj}kj=0, which reflects the memory effect of the non-local fractional derivative. Thus it requires on average O(NT) storage and the total computational cost is O(N2T) with NT the total number of time steps. This makes both the computation and memory expensive, specially in the case of long time integration. In order to overcome this difficulty, we propose a further approach to the fractional derivative. The idea consists in first splitting the convolution integral in (2.1) into a sum of history part and local part as follows:
Note that a comparable treatment is employed in reference [34]. Then the history part Ch(tk) can be rewritten as
hence we have
Using the simple recurrence relation (2.5), we define the discrete fractional differential operator Fαt by
It is not difficult to see that Fαthk=Lαthk for 1≤k≤NT. Comparing Lαthk in (2.4) with Fαthk in (2.6) and (2.7), the former requires all the previous time step values of h(t) while the latter only needs hk−1, hk and Fαthk−1. This implies that approximating CF0Dαthk by Fαthk considerably reduces the storage and computational costs as compared to Lαthk. Roughly speaking, replacing Lαthk by Fαthk allows to reduce the storage cost from O(NT) to O(1), and the computational cost from O(N2T) to O(NT).
Remark 2.1. The fast algorithm of Caputo derivative in [34] should be noted to retain an additional truncation error ε, whereas the fast algorithm of CF derivative does not introduce this error. Furthermore, it is worth mentioning that other algorithms, such as parallel computational methods [40], result in an augmented spatial complexity.
The following lemma provides an error bound for approximation Fαthk.
Lemma 2.1. Suppose that h(t)∈C2[0,T]. For any 0<α<1, let
Then
Proof. We consider proving the following estimate by mathematical induction:
First we have
where η1∈(0,t1). Therefore, (2.8) holds for j=1. Now suppose that (2.8) holds for j=k−1, we need to prove that it holds also for j=k. Similar to the proof of |R1|, we can easily get that
By combining (2.5) and (2.7), we obtain
The estimate (2.8) is proved. Hence
which prove the conclusion of the lemma. □
Remark 2.2. The second rate of convergence of L1 formula has been proven in [30] by different methods, here we obtained identical results herein. Note that the rate of convergence of L1 formula for classical Caputo fractional derivative with order α is 2−α, this result seems reasonable since Caputo-Fabrizio derivative has smooth kernel.
2.2. Discretization in time
We denote uk:=u(x,tk) and fk(x):=f(x,tk). Then from (2.6), (2.7) and Lemma 2.1, the time fractional derivative (1.5) at t=tk can be approximated by
where
Then Eq (1.1) can be rewritten as
where
with Cu,αi>0 for i=1,2,⋯,n. Notice that
we denote
the above equations are recast into
Let uk be the approximation for uk(x), and fk:=fk(x). Then the semi-discrete problem of Eq (1.1) can be written as
where
Moreover, by utilizing relation
we can easily derive an alternative formulation of (2.15)–(2.18) as follows
where
Remark 2.3. Since Lαit=Fαit, (2.19)–(2.22) can be also obtained by using Lαit in Eq (1.1). It is noteworthy that Eqs (2.15)–(2.18) offer computational advantages over Eqs (2.19)–(2.22). This is primarily attributed to the straightforward recurrence relation presented in Eqs (2.6) and (2.7). However, (2.19)–(2.22) is more appropriate for our analysis than (2.15)–(2.18), hence it play a crucial role in the subsequent sections.
Theorem 2.1. Let ˜Rk be defined by (2.12), then there exists a constant ˜c>0 such that
Proof. Without loss of generality, we assume that Δt∈(0,1). By the definition of Rk and the inequalities of (2.10), we have
On the other hand, since
for i=1,2,⋯,n, we get
This implies that
Therefore, there exists a constant ˜c>0 such that
which prove the conclusion of the lemma. □
Lemma 2.2. Let the coefficients b(αi)j,k be defined by (2.9), then for every i,
Proof. b(αi)j,k∈(0,1) can be easily obtained by the definition of σ(αi)j,k and the monotone property of the function g(x)=exp(x). Finally, note that
Using the above equalities and the fact exp(−αiΔt1−αi)∈(0,1) completes the proof of the lemma.□
Remark 2.4. (2.26) gives a easy way to compute all the coefficients b(αi)j,k.
Lemma 2.3. Let the coefficients ζj,k be defined by (2.23), then
Proof. By Lemma 2.2, and the definition of ζj,k, we can readily arrive at these conclusions.□
2.3. Stability and convergence analysis of the semi-discrete scheme
To discuss the stability and convergence of the semi-discrete scheme, we introduce functional spaces equipped with standard norms and inner products that will be utilized subsequently. Let L2(Λ) is the space of measurable functions whose square is Lebesgue integrable in Λ. Then
The inner products of L2(Λ) and H1(Λ) are defined, respectively, by
and the corresponding norms by
The norm ‖⋅‖m of the space Hm(Ω) (m∈N) is defined by
In this paper, instead of using the above standard H1-norm, we prefer to define ‖⋅‖1 by
It is widely acknowledged that the standard H1-norm and the norm defined by (2.27) are equivalent; therefore, we will adopt the latter in subsequent discussions.
The variational (weak) formulation of the Eqs (2.15) and (2.16)/(2.20), subject to the boundary condition (2.18), can be expressed as finding uk∈H10(Λ) such that for ∀v∈H10(Λ)
where u0:=u(x,0).
For the semi-discretized problems (2.28) and (2.29), we can establish a stability result as follows.
Theorem 2.2. The semi-discretized problems (2.28) and (2.29) is unconditionally stable in the sense that for all Δt>0, it holds
Proof. By mathematical induction. First of all, when k=1, we have
Notice that ‖v‖0≤‖v‖1, taking v=u1 and using the Cauchy-Schwarz inequality, we obtain immediately
Now, suppose
Taking v=uk in (2.29) gives
Hence, by using (2.30) and Lemma 2.3, we have
Thus, the proof is completed.□
Remark 2.5. In the proof of the following theorem, we will demonstrate that ζ−11,k is bounded. As shown in Eq (2.25), |κ|=O(1). Therefore, it follows that κζ−11,k is also bounded.
We now conduct an error analysis for the solution of the semi-discretized problem.
Theorem 2.3. Assuming n≠1 and 0<α1≤α2≤⋯≤αn−1≤αn<1. Let uk(x) be the exact solution of (1.1)–(1.3), {uk}NTk=0 be the solution of semi-discretized problems (2.28) and (2.29) with initial condition u0:=u0(x), then the following error estimates hold:
where the constant ˜c is defined in (2.24) and S is the length of Λ.
Proof. We shall establish the following estimate through a process of mathematical induction:
Let ˉek=uk(x)−uk, k=1,2,⋯,Nt. By combining (2.13) and (2.28), the error ˉe1 satisfies
Taking v=ˉe1 yields ‖ˉe1‖21≤‖˜R1‖0‖ˉe1‖0. This, in conjunction with (2.24), yields
Therefore, (2.32) holds for k=1. Assuming that (2.32) holds for all k=1,2,⋯,l−1, it is necessary to demonstrate its validity for k=l. By combining (2.13), (2.14) and (2.29), for ∀v∈H10(Λ), we have
Let v=ˉel in the above equation, then
Using the induction assumption and Lemma 2.3, we derive
Next we show that ζ−11,k is bounded. Considering that function h(α)=:−α1−α is decreasing on (0,1), we have
by combining Eq (2.26). Therefore,
Consequently we obtain, for all k such that kΔt≤T,
□
3.
Full discretization
3.1. A shifted Legendre collocation method in space
We shall begin by providing a comprehensive overview of fundamental definitions and properties pertaining to Legendre Gauss-type quadratures. Let PN(Λ) denote the space of algebraic polynomials of degree less than or equal to N with respect to variable x, and LN(x) be the Legendre polynomial of degree N on the interval [−1,1]. Then the discrete space, denoted by P0N(Λ):=PN(Λ)∩H10(Λ).
Let π1N be the H1-orthogonal projection operator from H10(Λ) into P0N(Λ), associated to the norm ‖⋅‖1 defined in (2.27), that is, for all ψ∈H10(Λ), define π1Nψ∈P0N(Λ), such that, ∀vN∈P0N(Λ),
From [41], the following estimate of projection holds:
Define the Legendre-Gauss-Lobatto nodes and weights as ξp and ωp, p=0,1,⋯,N, N≥1, where {ξk}Nk=0 are the zeroes of (1−x2)L′N(x), and
Moreover, the following quadrature holds
The discrete inner product and norm defined as follow, for any continuous functions ϕ,ψ∈C([−1,1]),
From [42], the discrete norm ‖⋅‖N is equivalent to the standard L2-norm in PN([−1,1]). If we denote {ˆξp}Np=0 and {ˆωp}Np=0 as the nodes and weights of shifted Legendre-Gauss-Lobatto quadratures on ¯Λ, then one can easily show that
Thus, we define the discrete inner product and norm on ¯Λ as follows
It is not difficult to obtain that
and
We introduce the operator of interpolation at the N+1 shifted Legendre-Gauss-Lobatto nodes, denoted by IN, i.e., ∀ψ∈C(¯Λ), INψ∈PN(Λ), such that
The interpolation error estimate (see [41]) is
Now consider the spectral discretization to the problems (2.28) and (2.29) as follows: find ukN∈P0N(Λ), such that
where
For {ujN}k−1j=0 given, the well-posedness of the problem (3.7) is guaranteed by the well-known Lax-Milgram Lemma.
3.2. Convergence analysis of the full discretization scheme
To simplify matters, we present the semi-discretized problems (2.28) and (2.29) in a compact form: find uk∈H10(Λ), such that
where
We denote by ‖⋅‖1,ˆN the norm associated to the bilinear form AˆN(⋅,⋅):
It follows from (3.3) that for all vN∈PN(Λ) the discrete norm ‖⋅‖1,ˆN is equivalent to the norm ‖⋅‖1 defined in (2.27).
Theorem 3.1. Assuming n≠1 and 0<α1≤α2≤⋯≤αn−1≤αn<1. Let {ukN}NTk=0 is the solution of the problem (3.7) with the initial condition u0N taken to be INu0(x), {uk}NTk=0 the solution of the semi-discretized problems (2.28) and (2.29). Suppose that uk∈Hm(Λ)∩H10(Λ) with m>1, for k=1,2,⋯,NT, then there exists a constant ¯c>0 such that
Proof. For any ∀vN−1∈P0N−1(Λ), denote ρkN:=ukN−vN−1. It is direct to check that
By virtue of (3.4) gives
hence
For the last term, by definition, we have
and
It is known that the following result holds (see e.g., [41,42]): ∀g∈Hm(Λ), m≥1, ∀δN∈PN(Λ),
Thus for ∀g∈Hm(Λ), m≥1, ∀gN,ρN∈PN(Λ) we have
Applying the above results to (3.9) and (3.10), we obtain
and
Let εjN:=uj−ujN, using (3.8) and the norm equivalence, for ∀vN−1∈P0N−1(Λ), we have
and
By triangular inequality
for ∀vN−1∈P0N−1(Λ), we obtain
and
The above estimate specially holds for vN−1=π1N−1uk, which implies
and
Similar to the proof of Theorem 2.2, we can immediately get the following conclusions:
Notice that
and the boundedness of κ and ζ−11,k, then there exists a constant ¯c>0 such that
□
4.
Numerical validation
4.1. Implementation
We provide a comprehensive account of the implementation of problem (3.7) using the shifted Legendre collocation method.
Considering problem (3.7), we express the function uk+1N in terms of the Lagrangian interpolants based on the shifted Legendre-Gauss-Lobatto points ˆξi,i=0,1,⋯,N, i.e.,
where cki:=ukN(ˆξi), unknowns of the discrete solution. hi(x) is the Lagrangian polynomials defined in Λ, which satisfies
where δij is the Kronecker symbols. Taking (4.1) into (3.7), and notice that the homogeneous Dirichlet boundary condition (1.3), then choosing each test function vN to be hl(x) (l=1,2,⋯,N−1), we have
Define the matrices
Then, we obtain the matrix representation of the above equation in the following form:
The linear system (4.2) can be solved in particular by the LU factorization or other related computational techniques.
Finally, we discuss about the calculation of Qk. When k=0, the initial condition u0N taken to be
It is not difficult to see that
which implies that u0N satisfies interpolation condition (3.5). When k≥1, suppose that
then
Furthermore,
In a word, we can easily obtain Qk at each iteration of time-step.
4.2. Numerical results
We present a series of numerical results to validate our theoretical propositions.
Firstly, to investigate the computational performance of two discrete fractional differential operators Lαt and Fαt, we test three examples from [30]. Denote β:=α1−α.
Example 4.1. Consider the function h(t)=tm (m≥1,m∈N), the Caputo-Fabrizio fractional derivative of order α with 0<α<1 of h(t) is written as
Example 4.2. Consider the function h(t)=cos(ωt), the Caputo-Fabrizio fractional derivative of order α with 0<α<1 of h(t) is written as
Example 4.3. Consider the function h(t)=exp(ωt), the Caputo-Fabrizio fractional derivative of order α with 0<α<1 of h(t) is written as
The proofs based on the method of integration by parts can be found in [8,30].
We choose m=4 in Example 4.1 and ω=5 in Examples 4.2 and 4.3, and set α=0.5, t∈[0,2]. Define the errors
for Lαt and Fαt operators, respectively, where NT is the last time step. Tables 1–3 give the numerical results of approximation error and CPU time with three examples. Here CPU time represents the total computation time, that is, the whole time for computing the approximations of Caputo-Fabrizio fractional derivatives at every time step. The convergence rates in Tables are given by
Tables 1–3 demonstrate that the errors of both Lαt and Fαt approximations are virtually identical, as a result of their equivalence (Fαthk=Lαthk). Moreover, both approximations have achieved second-order convergence of error, as stated in Lemma 2.1. However, we observe that the CPU time of Fαt approximation increases linearly with respect to NT, while the Lαt approximation increases almost quadratically. This suggests that the Fαt operator holds promise as it requires less storage and incurs lower computational costs than the Lαt operator during computation.
Secondly, we provide preliminary computational findings to demonstrate the efficacy of the finite difference/shifted Legendre collocation method (abbreviated as FCM).
Example 4.4. Consider the following three-term time-fractional diffusion equations:
where Λ=(0,π), 0<α1≤α2≤α3<1, and
with βi=αi1−αi, i=1,2,3. The exact solution of the Eq (4.3) is u(x,t)=(1+t2)sinx, which is sufficiently smooth. In our experiments, we set the parameters d1=1,d2=2,d3=3.
The following error norms have been used as the error indicator:
We test Example 4.4 with four cases:
Case 1. α1=α2=α3=0.5, then (4.3) reduces to the single-term time-fractional diffusion equation;
Case 2. α1=0.3, α2=0.5, α3=0.7;
Case 3. α1=0.2, α2=0.3, α3=0.4;
Case 4. α1=0.6, α2=0.7, α3=0.8.
In time discretization, we use Fαit operator. Table 4 shows the errors and temporal accuracy of FCM with polynomial degree N=20 at tk=1 for different cases of αi. Here convergence rates are given by
It can be observed that the FCM exhibits a second-order temporal convergence rate, which is consistent with our theoretical analysis.
Next, we check the spatial accuracy with respect to the polynomial degree N. In order to avoid the contamination of temporal error, we need fix the time step Δt sufficiently small. Here we take Δt=10−6, and terminate computing at tk=0.01 for saving time. Figure 1 shows the errors with respect to polynomial degree N at tk=0.01 in semi-log scale. Evidently, the spatial discretization exhibits exponential convergence as demonstrated by the nearly linear curves depicted in this figure. The aforementioned is known as spectral accuracy as expected since the exact solution is a sufficiently smooth function with respect to the space variable.
To further verify the numerical validity, we finally test a two-dimensional problem.
Example 4.5. Consider the following three-term time-fractional diffusion equations:
where Ω=[0,1]×[0,1], 0<α1≤α2≤α3≤1, and
The exact solution of the Eq (4.4) is u(x,y,t)=sin(2πxy)exp(t), which is sufficiently smooth. In our experiments, we set the parameters d1=1,d2=2,d3=3.
In this case, we denote {ˆξpq}Np,q=0:={(ˆξp,ˆξq)}Np,q=0 and {ˆωpq}Np,q=0:={ˆωpˆωq}Np,q=0 as the nodes and weights of shifted Legendre-Gauss-Lobatto quadratures on ¯Ω. Then we express the function uk+1N in terms of the two-dimensional Lagrangian interpolants based on the shifted Legendre-Gauss-Lobatto points ˆξij,i,j=0,1,⋯,N,
where ck+1ij:=uk+1N(ˆξi,ˆξj), unknowns of the discrete solution. hi(x) and hj(y) are the Lagrangian polynomials defined in Ix:=[0,1] and Iy:=[0,1], i.e.,
where δil and δjs are the Kronecker symbols. A linear system such as (4.2) can be readily derived. Here we take Δt=10−6. Figure 2 shows the errors with respect to polynomial degree N in semi-log scale. Thanks to the fast scheme (2.7), a small time step does not significantly escalate the computational burden in the time direction, thereby the proposed method is effective even for handling high-dimensional problems.
5.
Conclusions
In this work, we have developed a fully discrete scheme for the multi-term time-fractional diffusion equations with Caputo-Fabrizio derivatives. The proposed approach utilizes the finite difference method to approximate multi-term fractional derivatives in time and employs the Legendre spectral collocation method for spatial discretization. Specifically, we use the exponential property of Caputo-Fabrizio derivative to give a recursive difference calculation scheme, which offers benefits in terms of computational complexity and storage capacity. The proposed scheme has been proved to be unconditionally stable and convergent with order O((Δt)2+N−m). Numerical results show good agreement with the theoretical analysis. Due to its high resolution feature in spectral approximation, the proposed method can be extended to handle multi-term time-fractional diffusion equations in higher spatial dimensions.
Use of AI tools declaration
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
Acknowledgments
This work was supported by the Natural Science Foundation of Fujian Province (Grant No. 2022J05198), the Foundation of Fujian Provincial Department of Education (Grant No. JAT210293), and the Natural Science Foundation of Fujian University of Technology (Grant No. GY-Z21069).
We are thankful for the anonymous reviewers for their valuable comments and suggestions.
Conflict of interest
The authors declare no conflicts of interest in this manuscript.