1. Introduction
Diffusion processes hold a position of utmost significance in science. These processes, ranging from the spread of heat through materials to the dispersion of molecules in fluids, play a pivotal role in understanding various phenomena across diverse scientific disciplines. Traditionally, diffusion has been effectively described using classical, integer-order differential equations, such as Fick’s second law or the heat equation. However, these conventional models fail to capture the intricacies observed in systems characterized by anomalous behavior. For example, different aspects such as interactions between particles and memory effects have limited the classical approach in describing a large variety of experimental problems [
1].
One way to overcome the shortcomings of the classical approach involving integer-order differential equations is to instead use time-fractional diffusion equations, which have been found useful in many real-life processes where anomalous diffusion occurs [
1,
2,
3,
4,
5]. In the current paper, we consider the numerical solution of time-fractional diffusion equations corresponding to sub-diffusive models, where the order of the time-fractional differential operator belongs to
. Sub-diffusion refers to situations where particles spread more slowly than predicted by classical models [
1].
Since finding an exact solution to a time-fractional diffusion equation is not usually feasible in practice, the development of effective numerical methods is crucial for solving real-world diffusion problems. One of the more popular approaches is to use finite difference methods for the numerical solution of time-fractional diffusion equations (see [
6,
7,
8,
9,
10,
11,
12,
13,
14,
15] and the references therein). These methods usually discretize the spatial and temporal domains and approximate the derivatives using appropriate difference formulas. On the other hand, in the numerical solution of ordinary fractional differential equations, an often-used approach is to convert the fractional differential equation to a weakly singular integral equation and to solve the transformed equation using a collocation-type method (see, for example [
16,
17,
18,
19,
20]). However, only a few researchers have considered the numerical solution of time-fractional diffusion equations by collocation methods [
21,
22,
23,
24] (see also [
25]). In collocation methods one looks for an approximate solution in a finite-dimensional space and determines the approximate solution by requiring that it satisfies the equation on an appropriate finite set of points (on the so-called collocation points). If initial or boundary conditions are present, then the collocation solution will usually be required to fulfill these conditions, too. In particular, collocation methods that use polynomial splines and special non-uniform grids take into account the possible singular behavior of the exact solution of weakly singular integral equations. Moreover, these methods usually enable us to obtain a stable and high-order procedure with uniform convergence on the whole interval of integration. Therefore, in the present article, we are interested in developing a numerical scheme for time-fractional diffusion equations that combines space variable discretization and the classical piecewise polynomial collocation method on a non-uniform grid, where the grid points reflect the possible singular behavior of the underlying solution.
The space variable discretization (sometimes also called the method of lines [
26]) is a technique that involves discretizing with respect to the spatial variables and treating the resulting system as a set of coupled fractional ordinary differential equations. For our problem, this allows us to remodel the time-fractional diffusion equation into a system of ordinary fractional differential equations. To solve the obtained system of fractional differential equations, we reformulate it as a system of weakly singular Volterra integral equations of the second kind and employ a suitable collocation method for finding approximate solutions. Our approach enables us to construct a high-order numerical method for solving the sub-diffusion problem, despite the fact (see [
12,
27,
28]) that the temporal partial derivatives of solutions of time-fractional diffusion equations have, in general, weak singularities at the initial time
. It is worth mentioning that we can achieve a sufficiently high convergence order even when using polynomials of low degree.
The rest of the paper is organized as follows: In
Section 2, we present the underlying problem and necessary notations. Next, in
Section 3, we use the method of lines to create a system of ordinary fractional differential equations. In
Section 4, we reformulate this system as a system of integral equations and study the regularity properties of its exact solution. Then, we introduce a collocation-based method for finding approximate solutions in
Section 5 and study the convergence and convergence order of the proposed method in
Section 6. In
Section 7, we test our theoretical error estimates using some numerical experiments.
2. Time-Fractional Sub-Diffusion Model
Consider the time-fractional sub-diffusion equation
subject to the following boundary and initial conditions:
Here,
is the unknown function,
,
p is a positive constant (sometimes called the general diffusion coefficient),
with
,
and
, where
. Furthermore,
denotes the
-order Caputo fractional derivative of
with respect to the variable
t, which, for
, is defined [
29,
30] by
where
is the Euler gamma function:
For one-dimensional absolutely continuous functions,
y, on
, we will use the simplified notation,
, for the Caputo fractional derivative of
y:
Using
and
, with
,
, we denote the sets of
m times continuously differentiable functions on
and
, respectively; for
, we set
and
. In particular,
denotes the Banach space of continuous functions,
w:
, with the usual norm,
:
. Below, we will also use the following notations. For
the writing
; for
, we set
) means that
,
; note that
is a Banach space with respect to the norm
Using
, we denote the space of all essentially bounded measurable functions
, such that
where
means that the Lebesgue measure of the set
is equal to zero. For
, the writing
means that
,
, and
is a Banach space with respect to the norm
Finally, let X and Y be some linear spaces, and let be a given operator. Then, for a fixed and vector with , by we denote the vector , with .
Note that without the loss of generality, we can consider Equation (
1) only with homogeneous boundary conditions (
2) and (
3) since, in the case of more general boundary conditions
where
and
are some sufficiently smooth functions on
, Problem {(
1), (
4), (
6)} is easily transformed to a problem with homogeneous boundary conditions. Indeed, introducing (see, e.g., [
31]) an auxiliary function
by
the inhomogeneous boundary conditions (
6) for
u are transformed to the homogeneous boundary conditions for
v:
,
. Moreover, we see that both Equation (
1) and the initial condition (
4) maintain their original form with respect to the new unknown function
v:
where
and
where
For the existence and uniqueness of a classical solution
u to (
1)–(
4) (that is,
and
both exist in
Q and
u satisfies (
1)–(
4) pointwise), we refer the reader to [
31]. The regularity properties of solutions
u to (
1)–(
4) are described in [
12] (see also [
14,
28,
31,
32]). In particular, the smoothness of all the data of (
1)–(
4) does not imply the smoothness of the solution
u in the closed domain
, and the essential feature of all typical solutions to (
1)–(
4) is that the first-order derivative,
, in general, blows up as
(see [
12]). This is a significant obstacle for constructing high-order methods for the numerical solutions to (
1)–(
4).
On the other hand, in [
12], it is shown that when the data of Problem (
1)–(
4) has sufficient regularity, there exists a constant
such that for the spatial derivatives of the exact solution,
, to (
1)–(
4), we have
In our approach below, we assume that the solution,
u, to (
1)–(
4) satisfies the derivative bounds (
7). In particular, we will use this assumption already in the space variable discretization described in the next section.
3. Space Variable Discretization
We begin by developing a system of fractional differential equations from Problem (
1)–(
4) by space variable discretization using the idea of the method of lines. Let
,
. We introduce a uniform mesh on the interval
defined by
gridpoints
Using (
8) and a standard second-order difference formula
we approximate (
1)–(
4) using a system of equations
Here,
are the unknown functions, and
and
are defined by
Thus, we have for finding
a system of fractional differential equations in the form
subject to the initial conditions
where the function
is given by (
4), the functions
are defined by
and the constants
are determined by
For simplicity of presentation, we rewrite (
11) and (
12) in vector form
where
is unknown, its Caputo fractional derivative is defined componentwise by
, and
Below, we will also need the Riemann–Liouville integral operator,
, of order
, defined by
where
I is the identity mapping and
is the Euler gamma function. Note that the operator
is linear, bounded, and compact as an operator from
to
(see, e.g., [
33]). Moreover, we have for any
that (see, e.g., [
34])
4. Integral Equation Reformulation
Let
be an arbitrary continuous vector function, such that
. Then, with the help of (
5) and (
17), we obtain
where
is a constant vector and
It follows from (
16) that
in the form of (
18) satisfies the initial condition (
15) if and only if
Let now
be a solution to (
14) and (
15) such that
and
. Then, due to (
18) and (
19), we can rewrite Problem (
14)–(
15) in the form
where
Conversely, it is easy to see that if
is a solution to (
20), then
defined by (
18) with
is a solution to (
14) and (
15) and belongs to
. In this sense, Equation (
20) is equivalent to Problem (
14)–(
15).
In order to describe the existence, uniqueness, and regularity properties of the solution to Problem (
14)–(
15), we introduce weighted spaces,
and
, adaptions of a more general weighted space of functions introduced in [
35] (see also [
33]). For
, we define the weight function
and for given
,
,
,
,
, by
, we denote the Banach space of continuous functions,
, which are
q times continuously differentiable in
, such that
We see that if
, then for all
and
, the following estimation holds:
where
is a positive constant.
Further, given
,
,
and
,
, notation
means that
for
. The set
becomes a Banach space if it is equipped with the norm
Note that
Following the proof of Lemma 2.2, in [
33], we can prove the following result.
Lemma 1. Let and . Then operator , defined byis compact as an operator from to , thus also from to and from to . Furthermore, is compact as an operator from into , where . Theorem 1. Let and let , where , , . Let , . Then, the following statements are fulfilled.
- (i)
Problem (14)–(15) possesses a unique solution, , such that it and its Caputo derivative, , belong to . - (ii)
Let , , , with , given by (13). Then , the solution to Problem (14)–(15) and its Caputo derivative, , belong to , where
Proof. (i) We observe that
since
and
. Further,
is a compact operator from
to
since
is compact. Note that the homogeneous equation,
, has in
only a trivial solution,
. Therefore, using the Fredholm alternative, we obtain that the equation
possesses in
a unique solution
. Thus, Problem (
14)–(
15) has a unique solution
.
(ii) Let
,
,
. Then, clearly
. Since
, it follows from Lemma 1 that
is a compact operator from
to
. Therefore,
is also a compact operator from
to
. Since the homogeneous equation
has in
only a trivial solution, it follows from the Fredholm alternative that equation
has a unique solution
. Thus, Problem (
14)–(
15) possesses a unique solution
. □
5. Approximate Solutions for (14)–(15)
We construct an approximation
to
, the exact solution to Problem (
14)–(
15), as follows. First, we find an approximation
for
, the exact solution to (
20). Let
. We introduce on the interval
a graded grid,
, with the grid points
where
is the so-called grading exponent. We see that for
, the points (
24) are distributed uniformly, but for
they are more densely clustered near the left endpoint of the interval
.
For a given integer
, let
denote the set of polynomials of a degree not exceeding
k. We introduce the space of piecewise polynomial functions
where
is the restriction of function
to the subinterval
. Observe that the elements of the space
may have jump discontinuities at the interior points
of
.
Let
. Let
be a fixed system of collocation parameters satisfying
Using these collocation parameters, we introduce
m collocation points in each subinterval
by the formula
We find the approximation
for the exact solution
of equation
using collocation conditions
where
is defined by (
25) and
. If
, then, by
, we denote the right limit
. If
, then, by
, we denote the left limit
.
The collocation conditions (
26) with respect to
lead to a system of linear algebraic equations to find
,
, the exact form of which is determined by the choice of a basis in the space
. We can use Lagrange fundamental polynomial representation
where, for
and
, we set
if
and
Then,
and
for every
,
,
. Thus, we obtain a system of linear algebraic equations with respect to the coefficients
:
for
,
,
.
Having found
by the system (
28), we can determine
with the help of (
27). Thus, we obtain the approximation
to
, the solution to Problem (
14)–(
15), as follows:
6. Convergence Analysis
In this section, we study the convergence and convergence order of our method.
For given
, we define the interpolation operator,
, by
for any continuous function
. If
, then, by
, we denote the right limit
. If
, then
denotes the left limit
. Using operator
, the conditions (
26) for finding
with
,
, take the form
In order to prove Theorem 2 below, we need Lemmas 2–4. Lemmas 2 and 3 follow from the results of [
33,
35] and Lemma 4 follows from [
16].
Lemma 2. The operators, , , belong to the space andwith a positive constant c, which is independent of N. Moreover, for every , we have Lemma 3. Let be a linear compact operator. Then, Lemma 4. Let , , , , and . Let () be the Riemann–Liouville integral operator of order α. Assume that the collocation points (25) with grid points (24) and parameters satisfying are used. Moreover, assume that are such that a quadrature approximationwith appropriate weights is exact for all polynomials F of degree m. Thenwhere c is a positive constant independent of N and Theorem 2. Let the assumptions of Theorem 1 be fulfilled. Let , , and assume that the collocation points (25) with parameters satisfying and grid points (24) are used. Moreover, assume that parameters are chosen so that quadrature approximation (31) with appropriate weights is exact for all polynomials F of degree m. Then, the following statements are fulfilled. - (i)
Problem (14)–(15) possesses a unique solution, , such that . There exists an integer , such that for , Equation (30) possesses a unique solution , where for , determining by (29) a unique approximation to , the solution to (14) and (15), and - (ii)
If (0,b], with given by (13), then Problem (14)–(15) has a unique solution , such that and its Caputo derivative belong to and the following error estimate holds: Here, κ is given by Formula (23), r is a grading exponent given in (24), and is defined by (32).
Proof. (i) Existence and uniqueness are already proven in Theorem 1; thus, we only need to prove the convergence (
33). We note that
is a compact operator from
to
(see Lemma 1), thus also from
to
. Using the same proof idea as in Theorem 1, we can show that equation
possesses a unique solution
. In other words, operator
is invertible in
and its inverse is bounded:
. From this and Lemma 3, we obtain that for all sufficiently large
N, we can say that
Therefore, operator
is invertible in
for sufficiently large
N and
where
c is a constant independent of
N. Thus, for
, Equation (
30) has a unique solution,
, where
for every
. For
and
, the solution to equation
, we see that
Therefore, by (
35),
where
c is a positive constant independent of
N. It follows from (
18), (
29), (
36), and Lemma 1 that
where
is a positive constant independent of
N. Using Lemma 2, we see that convergence (
33) holds.
(ii) It follows from Theorem 1 part
(ii) (with
) that Problem (
14)–(
15) has a unique solution, such that
. From the proof of part
(i), we know that there exists an integer
, such that for
, Equation (
30) has a unique solution
, where for every
,
. Denote
With the help of (
30), we see that
, and therefore we obtain from (
37) the following equation with respect to
:
Since
, it follows from (
38) for every
that
We know from the proof of part (i) that
is invertible in
for sufficiently large
N and
for all
. Thus, there exists also the inverse of
in
for
and
Using (
39), (
40), (
35), and Lemma 2, we obtain
where
c is a positive constant independent of
N. From the definition of the operator
we see that
and therefore,
where
and
are some positive constants independent of
N. Due to
, we obtain
This leads to the estimate
where
is a constant that is independent of
N and where
and
are defined with the help of (
29) and (
18), respectively. Using Lemma 4, we see that the error estimate (
34) holds. □
In Theorem 3 below, we present the error estimate of our numerical method for solving Problem (
1)–(
4). We assume that the data of Problem (
1)–(
4) satisfies the conditions laid out in Theorem 2.1 in [
12]. Under these assumptions, it follows from [
12] that Problem (
1)–(
4) has a unique solution
u that satisfies (
1)–(
4) pointwise, and there exists a constant
C, such that
Theorem 3. Let the solution u to (1)–(4) satisfy the estimates (41) and (42). Let the assumptions of Theorem 2 be fulfilled. Then, the following error estimate holds:Here, (), is defined by (32), , κ is given by formula (23), r is a grading exponent given in (25), is given by (29), and c is a positive constant that is independent of n and N. Proof. It follows from (
9) and Theorem 2 that for
, we have
Note that
are fixed points defined by our numerical method, but
t belongs to
. □