1. Introduction
Stochastic time-fractional subdiffusion equations frequently arise when modeling heat conduction in materials with thermal memory under the influence of noise, as well as in the study of particle motion in viscoelastic fluids (see, e.g., [
1,
2,
3,
4]). To accurately describe random phenomena exhibiting long-range dependence, fractional noise becomes essential. Fractional noise is a zero-mean Gaussian process that is white in time, characterized by the Hurst index
. When
, it exhibits a positive correlation, reflecting a persistent autocorrelation structure. The increase in fractional Brownian motion (fBm) demonstrates this long-range dependence, meaning that the correlation between distant observations decays slowly over time, leading to persistent behavior. This property is particularly relevant for modeling natural phenomena where past events significantly influence future outcomes, such as in financial markets, climate studies, and fluid dynamics. Moreover,
implies a positive autocorrelation, indicating that an increase in one observation is likely to be followed by another increase, whereas a decrease is likely to be followed by another decrease. This positive correlation is often observed in systems with trends or cycles, making fBm with
a suitable choice for capturing such behavior [
5].
In this work, we consider a numerical method for solving the following stochastic semilinear subdiffusion problem that is driven by integrated fractional noise with
and
:
where
with the domain
and
is some regular domain with a smooth boundary. Here,
and
denote the left-sided Caputo fractional derivative and Riemann–Liouville fractional integral, respectively; see [
6,
7,
8]. Additionally,
is the fractional Brownian motion seen with the Hurst parameter
. The fractional noise
is a formal derivative of a centered Gaussian random field
, which is defined on a complete filtered probability space
; see
Section 2 for details. Here,
is the initial value and
is a nonlinear function that is specified in
Section 2.
In order to clarify the physical motivation of Model (
1), we briefly introduce one of its typical applications; see [
2] for details. Let the functions
,
, and
represent body temperature, energy, and flux density, respectively. In a homogeneous medium with constants
and
a, where
, we know that
which implies
But in some practical inhomogeneous medium environments, such as the heat conduction in materials with thermal memory subject to fractional noise, we need to express the energy term as
where
,
. Differentiating (
2), we obtain Model (
1); that is,
which describes the heat conduction in a non-homogeneous medium subject to Gaussian noise with the Hurst index
.
In recent years, significant advancements have been made in understanding the existence, uniqueness, and regularity of time-fractional stochastic partial differential equations. Key contributions include the works of Chen et al. [
1], Mijena and Nane [
3,
4], Liu et al. [
9], Anh et al. [
10], Kang et al. [
11], and Moulay et al. [
12]. Alongside these theoretical developments, a growing body of research has focused on using numerical methods to solve time-fractional stochastic partial differential equations. Jin et al. [
13] studied the strong and weak convergences of a fully discrete scheme to solve stochastic linear subdiffusion driven by additive noise using a semigroup approach. Gunzburger et al. [
14,
15] explored the time discretization and finite element method for approximating stochastic integral–differential equations driven by white noise. Wu et al. [
16] analyzed a fully discrete scheme and proved their error estimates for approximating linear stochastic subdiffusion problems driven by integrated white noise. Kang et al. [
11] examined the existence, uniqueness, and regularity of semilinear stochastic time–space-fractional partial differential equations. Hu et al. [
17] investigated the use of the L1 scheme and weak convergence for solving stochastic semilinear subdiffusion problems with additive noise. Giordano [
18] studied quasi-linear parabolic equations influenced by additive Gaussian noise that is white in time and fractional in space and characterized by a Hurst index of
. Cao et al. [
19] considered spatial semidiscretization for solving linear stochastic evolution equations driven by fractional noise with an
. Deng et al. [
20] explored a semidiscretization scheme for approximating semilinear stochastic wave equations driven by fractional noise with an
. Additionally, Deng et al. [
21] presented a unified convergence analysis for fractional diffusion equations driven by fractional Gaussian noise with an
. For numerical methods used for stochastic parabolic partial differential equations, we refer to Dai et al. [
22], Liu [
23], Yan [
24], Kruse [
25], Jentzen and Kloeden [
26], Chen et al. [
27], and the references therein.
Despite these advancements, relatively little attention has been paid to the theory and numerical methods used for solving stochastic semilinear subdiffusion equations driven by integrated fractional noise with an
. This problem presents two significant challenges: (1) the time-fractional derivative and time-fractional integration result in a solution operator lacking the semigroup property and (2) the interplay between nonlinear terms and fractional noise adds significant complexity to the numerical analysis.
In this work, we address these challenges by developing a numerical method for solving semilinear stochastic subdiffusion problems driven by integrated fractional noise. We first establish the time and space regularity of the mild solution using the semigroup approach and the properties of the Mittag-Leffler function.
We assume that the noise satisfies
where
is defined by [
13]
and
. Here,
denotes the space of the Hilbert–Schmidt operators introduced in
Section 2 below.
Let
be the solution to (
1). We obtain the following spatial regularity:
As for the temporal regularity, we denote it as follows:
and, with
, we obtain the following:
If
, then for
,
If
, then for
,
Let
denote the linear finite element space. The finite element method is used to find
such that with
,
where
is the
projection operator and
is the discrete Laplacian operator. Let
. Then, there is a positive constant
C such that the following conclusions hold:
For
and
, we obtain
where
.
For
, it holds that
where
and
.
Let
be a partition of
and
be the time step size. We shall approximate the time fractional derivative at
by using the L1 scheme, with
:
where
are defined as follows (see [
28]):
Further, we approximate the Riemann–Liouville fractional integral by using the first order convolution quadrature formula; that is,
where
are generated by
, as seen in [
29]:
As for the noise term
, we approximate it at
by using the Euler method, with
:
Let
denote the approximate solutions of
and
at
, respectively. We can therefore define the following fully discretized scheme:
where
and
.
Following the approach of Gunzburger et al. [
14], we introduce piecewise constant functions to handle the nonlinear terms and noise. This allows the fully discrete solution to be expressed as the convolution of a piecewise constant function using the inverse Laplace transform of a resolvent-related function, i.e.:
where the approximate operators
and
are defined in (
62) and (
63).
The optimal time discretization convergence order is established using the Laplace transform method and the resolvent estimates; thus, for
,
,
, and
, we obtain the following:
If
, then
where
.
Remark 1. Let
and
.
(i) When
, we obtainwhere
. (ii) When
, it holds thatwhere
and
. Remark 2. When
and
, we find, with
, the following:
(ii) When
which is consistent with the error estimate of [27]. Remark 3. When
and
, we find, with
, the following:
This paper is organized as follows:
Section 2 introduces some key notations and assumptions regarding the nonlinear term
f and the noise seen and establishes the spatial and temporal regularity of the mild solution.
Section 3 presents the finite element scheme and derives the error estimates for the semidiscrete case.
Section 4 extends the analysis to the fully discrete setting, proving the error estimates using the convolution representation of the fully discrete solution. Finally,
Section 5 provides some numerical simulations that validate and support our theoretical analysis.
We use C to denote a positive constant independent of the functions and parameters concerned, but it is not necessarily the same when it occurs in different places. We use c to denote a particular positive constant independent of the functions and parameters concerned.
2. Preliminaries and Notations
Let
be a real separable Hilbert space with the usual inner product
and norm
. Let
be a filtered probability space and
be the space of bounded linear operators from
to
. The fractional noise has the following Fourier expansion:
where
are real-valued mutually independent fractional Brownian motion (
) values with the Hurst parameter
, and
are the eigenpairs of the operator
, with
. The covariance operator
and
Q is a self-adjoint positive semidefinite operator. We use
to denote the space of Hilbert–Schmidt operators from
to
, equipped are with the inner product and norm:
where
denotes the Hilbert–Schmidt operator norm.
Let
be a Sobolev space defined by
with the norm
, where
are the eigenpairs of
, with
. For simplicity, we assume that
A and
Q have the same eigenfunctions
.
Assumption 1. When
, the operator A satisfies the following resolvent estimate:which implies that when
(see Yan et al. [28]), Assumption 2 ([
11])
. The nonlinear function f satisfies The conditions placed on f can ensure the existence and uniqueness of
.
Assumption 3 ([
13])
. Assumption 4. For
, we assumewhere κ is defined by (4), which ensures that
, where
is the Mittag-Leffler function defined in (14) below. Below we introduce some useful lemmas.
Lemma 1 ([
30,
31])
. For
and
, it holds that Lemma 2. For
and
, it holds that Proof. From the definition of
in (
9), we obtain
By applying Lemma 1 and the definition of the inner product in
, we obtain
which completes the proof of Lemma 2. □
Lemma 3 ([
32])
. Let
. Suppose that y is non-negative and satisfies the inequality where the function
and the constant
. Then, We denote
, and then (
1) can be formally written as
Taking the Laplace transform of (
13), we obtain
which implies that
Using the inverse Laplace transform, we obtain
where
Here,
increases from −∞ to ∞}.
According to the resolvent estimate (
11) and interpolation theory, we can easily show that
From (
18), and similar to the estimates of Lemma 4.1 in [
13], the above Mittag-Leffler functions have the following estimates when
and
:
Next, we shall consider the spatial and temporal regularities of the solution
. First we well show its spatial regularity.
Theorem 1. Let Assumptions 2–4 hold. Thus, we assume
when
. Then, there exists a positive constant C such that, with
, Proof. Simple calculation gives us
As for
I, using (
19) and with
,
, we find that
As for
, the estimate (
20), with
, and Assumption 2 lead to
where we require
.
When
, we obtain, from (
20),
, and the fact that
, the following:
Here,
comes from the time regularity in Theorem 2.
As for
, using Lemma 2, we obtain
From the symmetry of the integral region and integrand function, we find that
By using Assumption 4 about fractional noise, (
21) when
, and the variable transformation
, we obtain
Further, using Cauchy–Schwarz inequality, it holds that
Here, the definitions of
and
ensure the integrals
and
are finite.
The proof of Theorem 1 is complete. □
Next, we provide the temporal regularity of solution
.
Theorem 2. Let
. Let Assumptions 2–4 hold. We assume that
, as given by (5), and
. Then, for
and
, we know the following: (i) If
, then for
, (ii) If
, then for
, Proof. First, we divide
into three parts:
As for
I, from the fact that
and (
20) is true, we find that
By using Assumption 2, (
20),
, and
, we obtain
As for
, we find that
For
, by applying Lemma 2 and Assumption 4, we obtain
Case 1. If
, then by employing (
21),
, and the variable transformation
, we find that
where
. Further, by using the variable transformation
and the characteristics of the function
, when
, we obtain
Case 2. If
, then by employing (
21) and
, and similar to the discussion of case 1, we find that when
,
Thus, for
,
and for
,
Next, we turn to the estimation of
. Let
. We thus obtain
Note that for
,
where
.
Case 1. If
, i.e.,
, by choosing
, we obtain the following:
Case 2. If
, by choosing
and
, we derive the following:
Therefore, for
,
and, similarly, for
,
By combining the above estimates and using a continuous Gronwall inequality in Lemma 3, we complete the proof of Theorem 2. □
4. Time Discretization
In this section, we shall consider the time discretization scheme (
8). Since the homogeneous problem of Model (
1) has been studied in [
33], here we only need to discuss the numerical schemes and error estimates for the inhomogeneous problem of Model (
1), i.e.,
.
Taking the discrete Laplace transform of both sides of (
8), we obtain
We can write out the discrete Laplace transform of the sequences
as
respectively. We then arrive at
i.e.,
By using the inverse discrete Laplace transform, we find that
where
.
By denoting
, which are suitable approximations of
for
, respectively, we then obtain
We will show that
can be expressed as a convolution of the piecewise constant functions
and
. To obtain this, we first introduce the following piecewise constant functions,
, as defined by (
):
and
Similar to the proof of Lemma 2.1 in Wu et al. [
16],
in (
59) takes the following form:
Then,
can also be written as
Next, we introduce some lemmas which will be used in the error estimate of the time discretization.
Lemma 7 ([
13,
15])
. Let
and
be as defined above; then, we find obtainHere,
(similarly to
) means that there exist constants
and
such that Lemma 8. Denote
,
. Then, for
, we know that Proof. We only need to prove the second inequality as the first inequality is a special case of the second inequality. First, we know that
Next, we estimate
one by one. Note that by using (
11), we can easily obtain
For
, by using (
11), (
66), the mean value theorem, and Lemma 7, it holds that
Due to the fact that
, similar to the estimate of
, we derive
and
From (
67)–(
69), we obtain
□
Theorem 4. Assume that Assumptions 2–4 are satisfied and set
. Let
and
be the solution to (1) and (8), respectively. Let
be used for Theorem 3. Let
, and then we can obtain the following estimates for
,
,
, and
: Case 1. If
, thenwhere
. Case 2. If
, thenwhere
. Proof. The error estimate
has been proved in Theorem 3, so we only need to prove
.
By subtracting (
65) from (
41) when
, we find that
By simple calculation, we can derive that
Similarly, by using Lemma 8, we obtain
By applying Theorem 2, Lemma 7, and Assumption 2 we obtain
where
is the index of time regularity.
As for
J, we split this into three parts.
As for
, by applying Lemma 2, Assumption 4, and
, we find that
By using (
11) and letting
, we can obtain the following estimates for
:
We choose
, which ensures that
Case 2.
.
We note that
; thus,
where
.
Thus, with
, we obtain
Similarly, we also can prove this inequality with
:
Next, we estimate
. Let
. Following Lemma 2, Assumption 4, and Lemma 8, we find that, for
,
Case 1.
.
It holds that since
,
Case 2.
.
(ii) If
, we obtain
Combining Estimates (
78)–(
80), for
, we derive that
i.e.,
Finally we turn to estimate
. From the definition of
, it holds that
Further, by using Lemma 2, we find that
By using (
11), Lemma 7, and the fact that
, we can easily obtain
and, similarly,
Then, by substituting (
84) and (
85) into (
83), we obtain
Thus, from the estimate of (
77), we can obtain the estimate of (
86).
By combining the above estimates, we complete the proof of Theorem 4. □
By applying Theorem 4 and the fully discrete error estimate for the homogeneous equation corresponding to Model (
1) ([
13]), we obtain the following Theorem 5 for (
1).
Theorem 5. Let Assumptions 2–4 hold and
. Let
and
be the mild solution and fully discrete solution to Model (1), respectively. Let
. If we let
,
when
, then we have the following estimates for
,
,
,
: Case 1. If
, thenwhere
. Case 2. If
, thenwhere
. 5. Numerical Simulations
In this section, we will consider the numerical simulations required for solving (
1) when
,
:
where
and
where
represents fractional Brownian motion with the Hurst index
. It is straightforward to verify that
satisfies the nonlinear function Assumption 2.
Let the eigenfunctions of the operator
in
be given by
with the domain
. The eigenvalues are defined as
for
. If
, then
corresponds to white noise. If
, then
corresponds to smooth noise, since
.
Let
be a uniform partition of
with a time step size
. We define the following time discretization scheme for (
87) when
:
where the coefficients
are generated using the L1 scheme in (
6) and
are generated with Lubich’s convolution quadrature formula from (
7). Here,
is defined as
where
is generated using the MATLAB function
fbmid.m from MathWorks. Let
be a uniform partition of
with the space step size
h. We discretize the spatial variable using the linear finite element method to solve (
88).
In our numerical simulations, we set
and
. Since the exact solution of (
87) is not available, we compute a reference solution
using
. We then approximate the solutions
using the time steps
to calculate the time convergence orders.
According to Theorem 4, the time convergence order follows one of two cases, with
and
:
In particular, when
, i.e., the trace class case, and
, we obtain
For example, when
and
, the convergence order is
. When
and
, the convergence order is
. These results are verified in
Table 1. All computations were implemented in MATLAB R2024b on a laptop equipped with an Intel Core i5-8250U CPU (8th Generation, 1.6 GHz base frequency, up to 3.4 GHz), running on an Acer Aspire 5 model with a Windows 10 operating system.
When
, i.e., the trace class case, and
, we obtain
For example, when
,
, and
, the convergence order is
. When
,
, and
, the convergence order is
. These results can be observed in
Table 2.
In
Table 1, we set
and
. The table presents the average convergence orders for different values of
and
, confirming their agreement with our theoretical predictions.
In
Table 2, we set
and
, and a comparison of convergence orders for different
and
H again confirms their consistency with our theoretical expectations.
Conclusions
In this work, we introduced a new fully discrete scheme to approximate a stochastic semilinear fractional subdiffusion equation driven by integrated fractional Gaussian noise. The temporal and spatial regularity of the mild solution were proven using the semigroup approach. The finite element method was employed for spatial discretization, while the L1 scheme and Lubich’s first-order convolution quadrature formula were used to approximate the Caputo time-fractional derivative and the Riemann–Liouville time-fractional integral, respectively. The error estimates of the proposed scheme are proved in detail and explicitly show how the error depends on the parameters
,
, and H.