1. Introduction
With the rapid advancement of science and technology, anomalous diffusion phenomena observed in natural systems have attracted growing attention. It has become increasingly evident that classical differential equation models often fail to accurately capture certain diffusion processes. Over the last few decades, fractional diffusion equations have emerged as powerful modeling tools, finding widespread applications across diverse fields such as porous media, non-Newtonian fluid mechanics, and viscoelastic materials [
1]. In practical applications, crucial parameters such as boundary conditions, initial states, diffusion coefficients, and source terms are frequently unknown and thus must be determined using additional information, giving rise to inverse problems for fractional differential equations. Substantial progress has recently been made in the study of such inverse problems, including results related to uniqueness [
2,
3,
4,
5,
6,
7,
8] and numerical approaches [
9,
10].
The time–space fractional diffusion Equation (TSFDE), as a fundamental mathematical tool for characterizing anomalous diffusion, and it can be flexibly tailored to meet specific research objectives as follows: incorporating temporal correlations or memory effects yields time-fractional diffusion equations; accounting for spatial correlations or nonlocal effects leads to space-fractional diffusion equations; and considering both results in time–space fractional diffusion equations. Inverse problems for single-term cases of these equations have been extensively investigated both domestically and abroad. Notable studies have addressed backward problems [
11,
12,
13], source identification [
14,
15,
16,
17,
18,
19,
20,
21,
22], and the determination of the order of fractional derivatives [
23,
24,
25].
On the other hand, ultraslow diffusion characterized by logarithmic growth of mean squared displacement has been observed in various domains, such as polymer physics and particle motion in quenched random force fields. Such phenomena cannot be captured using classical advection–diffusion or conventional fractional diffusion models. To address these cases, distributed-order derivatives have been introduced, leading to distributed-order fractional diffusion equations. When the weight function of a distributed-order derivative is represented as a finite linear combination of Dirac delta functions with positive coefficients, the widely-employed multi-term fractional diffusion equation arises, which more faithfully models diffusion in systems with multiple time scales, coupled mechanisms, or nonlocal memory effects [
26].
In this study, we investigate an inverse problem associated with the following multi-term TSFDE:
where
denotes a bounded domain with a sufficiently smooth boundary
, and
d is the spatial dimension. For a fixed positive integer
m,
and
are positive constants such that
.
and
denote the spatially and temporally dependent source terms, respectively. The Caputo fractional left-sided derivative
is defined by the following:
where
denotes the Gamma function, and
represents the prescribed final time of the process under consideration.
The fractional orders
and the coefficients
are restricted in the admissible sets
with fixed
and
.
The fractional Laplacian operator, denoted as
, is defined for
within the range
, based on the spectral decomposition of the classical Laplace operator
. More specifically, in the context of a Hilbert space, this operator can be rigorously presented as follows:
and the operator
by
and it maps
onto
, In this expression,
and
represent the eigenvalues and the corresponding eigenfunctions of the Laplacian operator
, respectively. Here, the eigenfunctions
satisfy the equation
, subject to the given boundary conditions as follows:
therefore we set
For further discussion on Laplacian operator
, please refer to the literature [
27].
Suppose the unknown function
u is subject to the following initial and boundary conditions:
When the functions
,
, and
are all given, the problem consisting of Equations (
1)–(
3) is known as the direct or forward problem, where the goal is to compute the solution
. On the other hand, in the inverse problem, the aim is to recover the unknown time-dependent source function
. This identification is performed with the aid of supplementary measurement data at a fixed location
as follows:
where
denotes an interior observation point. The objective is to reconstruct
in the system (
1)–(
3) using the information provided by
.
Moreover, in the particular case where
, Equation (
1) reduces to the following single-term form:
Research into inverse problems for multi-term fractional diffusion equations has been mainly focused on time-fractional variants, yielding significant achievements in several directions as follows: identification of fractional derivative orders [
28,
29,
30]; reconstruction of various source terms [
31,
32,
33]; potential function determination [
34,
35,
36]; and inverse initial data problems [
37,
38], among others.
To the best of our knowledge, research focused on the inverse problem for multi-term time–space fractional differential Equations (TSFDEs) remains rather limited, Malik et al. [
39] investigated the identification of a time-dependent source and diffusion concentration within a TSFRDE characterized by multi-term Hilfer-type time derivatives and Caputo spatial derivatives, subject to homogeneous boundary conditions. Their work established the local Hadamard well-posedness of the inverse source problem, although no numerical algorithms were proposed. In contrast, our study emphasizes the question of uniqueness in recovering a time-dependent source term
from interior measurement data, thereby generalizing the single-term TSFDE scenario considered by Li et al. [
16]. As far as we are aware, our work is the first to tackle the inverse problem specifically for multi-term TSFDEs.
The rest of the paper is structured as follows. In
Section 2, we provide preliminary concepts and mathematical tools that serve as the foundation for the subsequent sections.
Section 3 details the application of the matrix transfer technique to the direct problem described by Equations (
1)–(
3), demonstrating the existence and uniqueness of strong solutions as well as developing an implicit finite difference scheme.
Section 4 delivers estimates on the stability and uniqueness properties of the inverse source problem. In
Section 5, the source identification problem is transformed into a variational problem using the Tikhonov regularization method, and then, an approximate solution to the inverse problem is obtained by employing the optimal perturbation algorithm. In
Section 6, numerical findings for six examples in one-dimensional and two-dimensional settings are examined. Finally, we offer a conclusion in
Section 7.
3. Regularity of the Solution and Difference Scheme for the Direct Problem
Let us denote the eigenvalues of the Laplacian operator under homogeneous Dirichlet boundary conditions by and the corresponding eigenfunctions by . Specifically, these eigenfunctions satisfy the spectral problem along with the boundary condition . Enumerating the eigenvalues according to their multiplicities, we have , and the family forms an orthonormal basis in .
Next, we introduce the definition of a strong solution for the direct problem (
1)–(
3), and we subsequently establish the existence and uniqueness of such a solution by following the approach in [
3].
Theorem 1. Suppose that , , and , and fix , . Then there exists a unique solution to (1)–(3), given bywhere and . Moreover, the following estimates hold:where and are positive constants that depend on , T, and Ω.
Proof. Based on Reference [
40] and the method of separation of variables, we can easily obtain the analytical expression
9 for the direct problem (
1)–(
3). It only remains to prove the regularity and estimates of the solution.
We first verify
and
. Define
then we have
. By using the result of Proposition 1, we get
and together with (
8), we can obtain
Therefore, we can obtain
where
represents positive constants depending on
,
T,
.
For
t,
, we have
From Proposition 1 and Lemma 1, we derive the following estimate:
we have
Consequently, we infer that
Since
for each
, by using the Lebesgue theorem, we can arrive at
which
.
Utilizing
and (
13), we deduce that
Next, we prove that
and
. Using (
9), we have
For
, by using (
7), we derive the estimate
and
Since
and
are uniformly convergent in
with respect to
t over any interval
for an arbitrary
, it follows in a manner analogous to the argument in the first part of the proof that
. Consequently, we obtain the estimate
where
is a positive constant depending on the parameters
, the final time
T, and the spatial domain
. □
In order to employ inversion techniques for solving the inverse problem at each step, it is first necessary to address the direct problem (
1)–(
3). In the subsequent analysis, we present an implicit finite difference scheme utilizing the matrix transfer technique, as detailed in [
41,
42,
43], to numerically solve the direct problem (
1)–(
3). For the sake of simplicity and clarity, our discussion is restricted to the one-dimensional multi-term TSFDE case.
In the finite difference framework, the time and spatial domains are discretized with step sizes and , respectively. The temporal grid points are given by for , while the spatial points are for . The notation represents the numerical approximation of the function u at the grid point .
We begin by considering the classical diffusion equation accompanied by the initial and boundary value conditions.
Introducing a finite difference approximation, we obtain
where
,
,
, and
h is the space step defined as
. The above equations can be expressed as the following system of ODEs:
where
and
,
,
Let
be a real, non-singular, symmetric matrix of size
. By virtue of its symmetry and non-singularity, there exists a non-singular matrix
such that
can be diagonalized as
where
is a diagonal matrix of the eigenvalues
of
.
With this preparation, using the methods described in References [
41,
42], we now reformulate the direct problem (
1)–(
3) in the following matrix representation:
where
,
. The time-fractional derivative is approximated by
where
, this scheme was used in [
44]. We have
, and
Subsequently, the implicit difference scheme defined by Equations (
16) and (
17) can be recast in the following matrix form:
where
and
The above describes a finite difference discretization of a one-dimensional multi-term space–time fractional diffusion equation. In the two-dimensional case, the approach is analogous; however, the matrix obtained from the spatial fractional-order discretization no longer maintains a symmetric tridiagonal form. Instead, it takes on a symmetric tridiagonal block structure as follows:
where
The detailed discretization procedure for the two-dimensional scenario will not be reiterated here.
5. The Inversion Algorithm
In the preceding section, we provided a theoretical justification demonstrating that the source term
can be uniquely identified from the measurement data collected at a single boundary point. Building upon this result, in the present section, we propose a computational approach to obtain an approximate solution for the source term. This method leverages the additional data specified in (
4) to enhance the numerical reconstruction of
.
Through the integration of Equations (
4) and (
9), a first-kind Volterra integral equation is derived as follows:
where the kernel function takes the form
and the right-hand side term
represents the measured data at the interior point
as follows:
where
and
.
As established in the seminal works of [
16,
45], the kernel function
exhibits weak singularity over the domain
, and the associated operator
is compact when mapping from
to
. Consequently, the inverse source problem (
1)–(
4) is inherently ill-posed. This mathematical characteristic implies that the solution
lacks continuous dependence on the input data, meaning that even infinitesimal perturbations in the measurement data may lead to significant deviations in the reconstructed solution.
To resolve this challenging inverse problem, we make use of the optimal perturbation algorithm to estimate the time-dependent source term
. Let us consider
as a complete set of basis functions. The source term can then be approximated through the following finite-dimensional expansion:
where
represents the
K-th order approximation to
,
denotes the truncation level controlling the approximation accuracy, and
are the coefficients of expansion.
We define the finite-dimensional approximation space as follows:
and associate each approximation
with its coefficient vector
. This parameterization transforms the infinite-dimensional inverse problem into a finite-dimensional optimization problem.
To handle the intrinsic ill-posedness of the inverse problem and to ensure numerical robustness, the Tikhonov regularization method is employed in solving the following minimization problem as follows:
where
is taken as the regularization parameter that controls the trade-off between solution accuracy and stability;
is the solution of the direct problem (
1)–(
3) for any prescribed
, given by (
26); and
denotes the measured data at the observation point
. The first term in the functional
measures the discrepancy between the computed solution and the observed data, while the second term serves as a regularization term that penalizes large values of the coefficient vector
, thereby ensuring the stability of the numerical solution.
To numerically solve the inverse problem (
27) for reconstructing the source term
, we employ an optimal perturbation algorithm. For any given coefficient vector
, we consider the following iterative update scheme:
where
represents the perturbation term at the
j-th iteration. For notational simplicity, we will hereafter denote
and
as
and
, respectively.
Applying the first-order Taylor expansion to
about
while neglecting higher-order terms yields the following linear approximation:
To determine the optimal perturbation, we minimize the following regularized error functional:
where
serves as a regularization parameter with the following adaptive selection strategy:
where
n denotes the iteration count,
represents a predetermined threshold (set to 5 in our implementation), and
(chosen as 0.8) controls the transition rate of this sigmoid-type regularization parameter, following the approach in [
46].
We now proceed to discretize the temporal domain
using a uniform grid,
, where
S denotes the number of grids. The above
norm can then be approximated by the discrete Euclidean norm, yielding the following regularized minimization problem:
where the sensitivity matrix
is defined component-wise as follows:
where
denotes the finite difference step size. The vectors
and
represent the computed and measured data, respectively, as follows:
Following standard regularization theory [
47], the minimization of (
30) leads to the following normal equation:
The optimal perturbation is consequently obtained through the following regularized solution:
The iterative inversion process continues until either the maximum iteration count is reached or the perturbation satisfies the convergence criterion as follows:
where
is a prescribed tolerance threshold governing the termination of the algorithm.
6. Numerical Experiments
In this part, we prove the effectiveness of the optimal perturbation algorithm with numerical results by using five examples for the one-dimensional scenarios and two-dimensional scenarios. The algorithms convergence and stability are examined.
We set
in all our experiments. The accurate data are perturbed to create the noisy data randomly, i.e.,
with the use of
, the corresponding noise level can be calculated.
Intending to show the accuracy of the numerical solution, the relative root mean square (RRMS) error is estimated to be
where
n denotes the total number of uniformly distributed points on the time interval
, the term
represents the reconstructed source term at the final iteration, while
corresponds to the exact (analytical) solution.
For generality and simplicity in numerical experiments, unless otherwise specified, we adopt the following parameter settings: the order of and the regularization coefficients are all set to 1. The convergence threshold is chosen as the first iteration , ensuring high precision in the iterative process. The initial guess for the first iteration is a zero vector (i.e., ), and the numerical differentiation step size is assigned a value of 0.01 to balance computational accuracy and efficiency.
6.1. One-Dimensional Case
To maintain generality, we discretize both time intervals and , using 51 uniformly distributed grid points when addressing the direct problem. For the one-dimensional scenario, the spatial domain is specified as . Throughout our computations, unless otherwise indicated, the observation location is chosen as . The following three examples are explored via numerical experiments to evaluate the effectiveness of the proposed method.
Example 1. Suppose the initial condition is , and the spatial source function is , with fractional orders , and temporal source component . To simulate the measured data, we solve the direct problem (1)–(3) using the finite difference method. By setting and choosing the basis functions , we generate the required input data at the observation point . The numerical results for different levels of noise, specifically
and
, are illustrated in
Figure 1. The results clearly demonstrate that the proposed approach remains both robust and accurate even when the observation data are corrupted by noise up to
.
We fixed the relative noise level at
and the fractional order at
.
Table 1 presents the RRMS errors
as defined in (
34). For Example 1, different values of
are presented in
Table 1. The data clearly demonstrate that as the components of
decrease, the numerical precision is enhanced, indicating the improved accuracy of the method. Additionally, for Example 1 with various settings of
, the RRMS errors
are reported in
Table 2, where the noise level is maintained at
and the parameters are set to
,
, and
. The results illustrate that higher values of
are associated with increased numerical accuracy. Furthermore, by fixing both the relative noise level at
and the parameters
,
,
, and
, the influence of the observation point
on the RRMS error
is summarized in
Table 3. These results indicate that the accuracy of the solution exhibits some dependency on the specific choice of
, suggesting there is a certain sensitivity pertaining to the method to the selection of this parameter.
Figure 2 presents the numerical results for Example 1 under various values of
and for several noise levels,
and
, with the parameters fixed as
;
…; and
. The computed solutions exhibit a strong agreement with the true function
, which demonstrates the high accuracy of the numerical approximation. These findings indicate that the optimal perturbation algorithm is both effective and robust when identifying the source term, even in the presence of noise.
Table 4 displays the RRMS errors
for Example 1 across different values of
, under a fixed relative noise level
and
;
, …; and
. As shown, the RRMS error
changes only marginally as
varies, indicating that the method maintains stable accuracy with respect to parameter
.
Table 5 reports the RRMS errors
for Example 1 at different locations of
, fixing the noise level at
, with
;
, …;
; and
. The results show a slight dependence on the value of
, but overall the reconstruction remains stable, further supporting the robustness of the approach.
Example 2. Assume the initial data are given by and let the coefficient and source functions can be defined as , with parameters , and , . To generate the input data at the observation point , the direct problem (1)–(3) is solved numerically using the finite difference method. We set and . Figure 3 presents numerical results for varying values of
and for different noise levels, namely,
and
. It can be observed that the numerical approximation accurately captures the steady-state boundary for all tested conditions. However, numerical accuracy begins to degrade toward the initial time, specifically at
.
Example 3. We consider a non-smooth scenario characterized by a cusp in the source term, specifically choosing . The initial data are set as , while the space-dependent coefficient is chosen as , With parameters fixed at , , and , the direct problem (1)–(3) is solved using the finite difference method to obtain the measurement data employed in the reconstruction. For the numerical experiments, we select and use the basis set . Figure 4 displays the computed results for different values of
with noise levels of
and
. The results show that, overall, the identification quality remains satisfactory, except in the vicinity of
and near the location of the non-smooth point, where some loss of accuracy is observed due to the underlying cusp in the source term.
6.2. Two-Dimensional Case
The spatial coordinates are assigned as . In the case of a two-dimensional case, boundary is applied and the space domain is supposed to be . The grid points on are , and we calculate 50 layers for a given step at time direction when solving the direct problem. We provide the numerical outcomes for three examples to show the precision and consistency of our suggested approach.
Example 4. Given a source function , , and the initial data . The measurement data are obtained by solving the direct problem (1)–(3) using the precise input data from the finite difference approach, where we take , , and and set ,. Example 5. The time-dependent source term is taken into consideration. Consider the initial data and take a source function . We construct the input data on observation point by solving the direct problem (1)–(3) by using a finite difference method and set , . Example 6. We take into account a time-dependent source term . Let the initial data and choose a source function . We create the input data on observation point by utilizing the finite difference approach to solve the direct problems (1)–(3) and set , . Figure 5,
Figure 6 and
Figure 7 show us the numerical outcomes for Examples 4–6 at varying noise levels in the situations of
, and
. It can be seen that our suggested approach works well for solving two-dimensional instances as well.
7. Conclusions
In this study, we explored the identification of time-dependent source terms within multi-term TSFDE. Initially, we rigorously established the existence and uniqueness of solutions to the associated direct problem, ensuring a solid theoretical foundation. Additionally, we derived both the uniqueness and stability estimates for the corresponding inverse problem, which seeks to recover the time-dependent source term using supplementary internal observation data.
From a computational perspective, the direct problem was efficiently solved via an implicit finite difference scheme augmented by the matrix transfer technique, allowing us to address the challenges posed by fractional derivatives. For the inverse problem, we recast the source identification task as a variational problem utilizing Tikhonov regularization, and we employed the optimal perturbation approach to obtain stable approximate solutions. Comprehensive numerical experiments, encompassing five one- and two-dimensional test cases, verified the proposed algorithm’s effectiveness and robustness in identifying time-dependent source terms for multi-term TSFDEs, delivering high accuracy even in the presence of noise.
However, when using the optimal perturbation algorithm to obtain approximate solutions for inverse problems, the parameter K and the basis functions have no fixed selection method and are generally chosen empirically, which poses significant limitations in practical applications. In the future, we plan to explore algorithms such as the conjugate gradient method, trust region methods, or deep learning, which can avoid the need to select basis functions and the number of terms in the expansion, thus making the approach more suitable for practical applications. On the other hand, future research will focus on extending this method to the identification of source terms in multi-term time–space fractional diffusion-wave equations, thereby further broadening its scope of applications.