1. Introduction
The Sturm–Liouville equation appears in certain practical areas, such as heat flow and vibration problems, electroencephalography applications, and other areas of physics. It has a relevant role in quantum mechanics, and some of these problems are formulated in unbounded intervals. On occasion, these problems are described by differential equations with highly oscillatory coefficients. A particular characteristic of these coefficients is that they are not square Lebesgue integrable. The study of differential equations involving integrable Henstock–Kurzweil functions has been developed by several authors, for example, [
1,
2,
3,
4,
5,
6,
7]. In [
8], Pérez et al. introduced the KH–Sobolev space on bounded intervals and guaranteed the existence and uniqueness of the solution to some boundary value problems involving Kurzweil–Henstock integrable functions on
. In this paper, particularly in
Section 3, we introduce the KH–Sobolev space for unbounded intervals, and then we apply these spaces and the Fredholm alternative theorem to establish the existence and uniqueness of the solution to the Sturm–Liouville differential equation
subject to the Dirichlet boundary value conditions
where the derivative is in the weak sense,
q is a Lebesgue integrable function,
f is Kurzweil–Henstock integrable,
is a function of bounded variation on all compact intervals
,
is Lebesgue integrable, and
on
. The solution is proven to be stable under small variations of
f. See
Section 4.
The Finite Element method (FEM) has been used to give approximations of the solution of a differential equation when the functions involved are continuous or square integrable. León–Velasco et al. in [
9] used the FEM to find numerical approximations of the solution of certain Sturm–Liouville-type differential equations involving Henstock–Kurzweil integrable functions. The existence and uniqueness of the problems given in [
9], as well as the convergence of the FEM are not studied in that paper. In
Section 5, we give conditions for the existence and uniqueness of elliptic problems on a bounded interval. In
Section 6, we show a scheme for the convergence of the Finite Element method.
2. Preliminaries
Throughout this paper,
I will be an interval of the form
,
,
, or
. Positive functions
defined on
I will be called gauge functions. These functions will control the refinements of the partitions in the Kurzweil–Henstock integral. Next, we give the definition of this integral when
I is a bounded interval. Let
be a gauge function. A tagged partition
of
I is said to be
-fine, when for every
,
Definition 1. A function is said to be Kurzweil–Henstock integrable (in abbreviation, KH-integrable) if there exists a number with the property that for every , there exists a gauge such that for every tagged partition of , if Pfine, then The Kurzweil–Henstock integral of f on is denoted and defined by
Now we define the Kurzweil–Henstock integral for non-bounded intervals. Let and be a gauge function. A tagged partition of is said to be -fine, if , and for each , .
Definition 2. A function f defined on is said to be Kurzweil–Henstock integrable (KH-integrable) if there exists a number with the property that for every , there exists a gauge such thatfor all tagged partitions P of which is fine. For functions defined over the intervals and , we have similar definitions. In the case where f is defined on , we can extend f on , assuming that . In this situation, we say that f is Kurzweil–Henstock integrable on if f extended to is -integrable. Similar considerations are given for intervals of the form , .
The space of functions which are Kurzweil–Henstock integrable on
I is denoted by
. The Alexiewicz seminorm for this space is denoted and defined as
The Lebesgue space
, for
, is defined as the set of Lebesgue-measurable functions
f on
I for which
. The seminorm of this space is given by
It is well-known that
. See ([
10], Corollary 4.80). This inclusion is strict, see, for example, ([
11], Example 3.12). In particular, if
, then
. Moreover, when
I is a bounded interval, it follows that
. Unfortunately, the space
is not a complete space.
The variation of a function h on the interval I is denoted by . If , then h is of a bounded variation on I, and we write . The functions of bounded variations are the multipliers of the KH-integrable functions. This allows the following Hölder-type theorem to be established:
Theorem 1. ([12], Lemma 24) If and , then and A function F is on a set E, if for every there exists such that for every collection of non-overlapping closed intervals with endpoints in E, it follows that if , then . Additionally, the function F is on a bounded interval J, if F is continuous on J and , where is a sequence of subsets of J and F is on each . In the case of unbounded intervals, we say that f is on if f is on each compact interval and f is continuous at ∞. For other intervals, we have similar definitions.
Theorem 2. ([13]) Let be the left end of the interval I. The following properties hold: - 1.
If and , then is on I and a.e. on I. Moreover, if f is continuous at then .
- 2.
F is on I if, and only if exists a.e. on I and for all .
Theorem 3. ([12], Lemma 25) Let and be functions such that and for each compact interval , - 1.
exists,
- 2.
there exists such that for all .
Thenfor all with . Moreover, ifexists, then 3. The Kurzweil–Henstock-Sobolev Space for Unbounded Intervals
Let be the space of functions for which there exists a partition of such that for every , , and , , , ⋯, , , exist, for . Now, we define the space for an interval I as follows:
, when ,
there exists such that supp and , when ,
there exists such that supp and , when ,
there exists such that supp and , when .
It is clear that if , then v and belong to . Throughout this section, we will only consider the interval . We denote by the space of functions defined on that are -integrable on every compact interval.
Lemma 1. Let and suppose that for all . Then the function f is zero a.e. on .
Proof. Let
. Then we show that
for all
. Let
, choose
such that
, and define the function
From ([
14], Theorem 12.5), there exist
and
such that
and
The case when is proved in a similar way. Consequently, . On the other hand, a.e. on , thus, a.e. on . ☐
Corollary 1. Let and suppose that for all . Then there is such that a.e. on .
Proof. Take
satisfying
. Let
and define
Then,
and so
where
are such that
. Let
and
, then
and for every compact interval
J,
and
for all
. Therefore, by Theorem 3,
Therefore, by Lemma 1,
a.e. on
. ☐
Lemma 2. Let and define Then, andfor all . Proof. Then, for every compact interval
,
Thus,
exists; moreover,
for all
. Therefore, by Theorem 3,
From (
2),
for all
. Therefore, by Hake’s Theorem,
Now, since
and
, it follows by the Dominated Convergence Theorem that
☐
Definition 3. The KH–Sobolev space is defined as The weak derivative of is denoted and defined by where w is as in Definition 3. Lemma 1 implies the uniqueness of , except for sets of measure zero. It is clear that if and is a scalar, then and .
Remark 1. 1. If and , then by Lemma 2 and .
- 2 .
If a.e. on and is on , then and a.e. on .
Theorem 4. For each , there is for which a.e. on and Proof. We set
. From Lemma 2, we obtain that
for every
, and since
,
for all
.
Consequently, for each
,
Therefore, by Corollary 1, there is for which a.e. on . Putting , we obtain the conclusion of the theorem. ☐
By ([
6], Corollary 2.4) we have the following integration by a parts formula for the weak derivative.
Theorem 5. If are in , then is also in and Moreover, for every , if the product is in and , , and , then 4. Sturm–Liouville Differential Equations for Unbounded Intervals
We denote by the space of functions such that , and by the space of functions such that , and , where is the left end of the interval I and is the right end.
Observe that if , then by Theorem 4, there is so that a.e. on I, therefore . In this way, we can equip the space with the seminorm .
In this section, we will only consider ; however, the results are also true for intervals of the form or .
Problem 1. Let , and ρ be a function such that for all compact intervals , and . We solve the following boundary value problem:
Find that satisfies This problem is equivalent to the following variational problem:
Find
such that
We will provide a solution to this variational problem. Define
where
Additionally, for every
, define
and
where
Then,
. Define the operators
and
by
Then,
is a bilinear operator and
is a linear operator that satisfies
for all
. Therefore, equality (
5) is represented by the equation
If there were
such that
then the equality
would be satisfied, for every
. Therefore,
u would be a solution to the variational problem (
5). We will show, using Fredholm’s alternative theorem, that under certain conditions there is indeed a solution to Equation (
10).
Theorem 6. Suppose that Y is a compact Hausdorff topological space. A subset of is relatively compact in the topology induced by the uniform norm if, and only if:
- (i)
, for all .
- (ii)
For every and , there exists a neighborhood of such thatfor all and .
Theorem 7. The operator is compact.
Proof. Let
be such that
for all
and some
. Consider the set
We use Theorem 6 in order to prove that
is relatively compact in
. Note
where
.
- (i)
Let
and
. Then
- (ii)
Let
and
. We suppose that
. Since
is continuous at
and
, it follows that similarly to ([
15], Lemma 4.1) there is a number
such that for every
,
Let
and
. Then,
Suppose now that
. Since
and
are continuous at
s, it follows that there is a number
which satisfies that if
, then
Let
and
. We suppose without loss of generality that
. Then,
From Theorem 6, we have that
is a compact set in
. Therefore, we have proved that
is compact when
E is a bounded subset of
. Take a bounded sequence
in
. Then,
is a compact set in
; consequently, there is a subsequence
of
and
such that
uniformly. Since
and
it follows by the Dominated Convergence Theorem that
and
We set
. Then,
,
and
Therefore,
. Finally,
Thus, is a compact operator. ☐
Theorem 8 (Fredholm’s alternative theorem). Let X be a normed space and consider a compact linear operator . Then, the transformation is injective if only if is surjective. Therefore, is bounded, when is injective.
Proposition 1. If the homogeneous problemhas only the trivial solution, then the operator is injective. Proof. Take
u in the kernel of
. Then,
a.e. on
. Since
is
on
, it follows by Remark 1 that
a.e on
, thus
a.e. on
. Since
is
on
, it follows again by Remark 1 that
and
a.e on
. Therefore,
u is a solution of the homogeneous problem. Consequently,
a.e on
. ☐
Proposition 2. If ρ and q are positive, then the homogeneous problemonly has the trivial solution. Proof. Let
be a solution of the homogeneous problem. Then
By Theorem 4, there exists
such that
a.e. on
and
for all
. From Remark 1,
and
. Then by (
12),
a.e. on
. Now, observe that
,
,
and
. Thus by Theorem 5, we have that
which implies that
Since and , we have that a.e. on . ☐
Theorem 9. Let , and be such that for all compact interval , and is on . Suppose that either of the two following conditions hold:
- (i)
The functions are positive;
- (ii)
The function zero is the unique solution of the homogeneous problem (11).
Then the following problemhas a unique solution u in the space and the solution u depends continuously on the data f. Proof. By Theorem 7, the operator
is compact, and by Proposition 1,
is injective. Thus, by Fredholm’s alternative theorem, the transformation
is surjective. Then, there exists
such that
. Hence,
for all
, consequently,
u is a solution to the variational problem (
5), and so
u is a solution to the boundary problem (
4). It is clear that if
w is another solution to the boundary problem (
4), then
is a solution the homogeneous problem, which implies that
a.e. on
.
Now, let
be a sequence in
such that
, and for each
, let
be a solution to the boundary problem
On the other hand, by Fredholm’s alternative theorem,
is a bounded operator. Then
Consequently, when . ☐
Example 1. In order to derive the steady-state heat conduction model, consider a non-uniform bar of infinite length with cross-sectional area Λ. Let be the temperature, the heat flux and the source term that models the generation or loss of heat at each point of the cross-section of the bar at position t, where . If is a small and arbitrary portion of the bar, then by the law of conservation of energy, we have, Dividing by and taking the limit as , we have If is the thermal conductivity of the bar, then by Fourier’s heat conduction law, , we obtain the steady-state heat conduction equation As a particular example, let us consider a non-uniform bar such that its property of conducting heat is greater as the position increases, then the thermal conductivity can be modeled by the function Furthermore, if we assume that heat is continuously lost in certain portions, and in others it is gained due to some source with null effect at distant locations, then one way to represent this behavior is by the function Setting the boundary conditions y , we obtain the problem with boundary values for the temperature : As , ρ is a function of bounded variation on every compact interval , and the function , and it follows by Theorem 9 that the problem (14) has a unique solution. 5. Sturm–Liouville Differential Equations for Bounded Intervals
Let us begin this section by showing that when
I is a compact interval on
we have
where
and
is the interior of
I. Let
a be the left end of the interval
I and
b be the right end, and take
, then there exists
such that
for all
. Take an arbitrary
, then
. Since
is dense in
, there exists a sequence
in
such that
. Consider
such that
and define the functions
by
where
, and
Then is a sequence in . We prove that
.
For the first convergence, observe that
Now, since
it follows that
Thus, (
16) tends to zero.
The second convergence is deduced by the following. From Theorem 1,
Finally, since each
, it follows that
therefore
However, it is also true that
Consequently, due to the uniqueness of limits, it follows that
Therefore, .
From (
15), we obtain the following sequence of inclusions:
As in
Section 4, we will consider
with the seminorm
. The form of this semi-norm is required in the following section. Based on ([
8], Theorem 4.3) and the results of the previous section, we state the following theorems:
Theorem 10. Let , and be such that is on . Then, the following assertions are equivalent:
- 1.
The following boundary problemhas a unique solution in . - 2.
The following variational problemhas a unique solution in .
Theorem 11. Let , and be such that is on . Suppose that either of the two following conditions hold:
- (i)
The functions are positive;
- (ii)
The function zero is the unique solution of the homogeneous problem:
Then the following problemhas a unique solution u in the space , and there is for whichwhere is arbitrary and is the solution of the problem: 6. Finite Element Method
In this section, we give a finite element method scheme for
-integrable functions. We consider
and
f as in
Section 5. Let
be the unique solution to the boundary problem
Then by Theorem 10,
and
Let
and
be a partition of
. We set
and we consider the finite element space
given by
Let
be an interpolate of
f on
, that is,
for all
. Then from Theorem 11, there exists
such that
Now, we will find
such that they satisfy
A basis for the space
is given by the functions
,
, defined as
Then, for every
,
Observe that if
are defined by
then the Equations (
20) and (21) are equivalent to
Let
and define
Then
M is symmetric and tridiagonal. Additionally,
M is positive-definitive, because if
then
If
then
, therefore
, which is a contradiction. Thus,
. Consequently,
M is invertible. Thus, there exist
and
unique solutions of the systems
and
, where
Consequently,
and
satisfy (
24) and (
25) for all
.
We will now estimate the error committed. First, observe that for every
and
, if
then
and
Indeed, equality (
27) is deduced by the following
Equality (
28) is obtained from
and equality (
29) is deduced by
From (
17) we have that
. Then by (
18) and (
20), it follows that
Observe that
is the optimal approximation for
u, that is,
where
for all
. Indeed, by (
29),
This implies that
and so the inequality (
31) is satisfied. Now, take an interpolate
of
u on
, that is,
for all
. We take
such that
, for all
. Then, for every
, there exists
between
s and
such that
for all
. Consequently,
On the other hand, from (
18)–(21) we have that
Thus by (
27),
but, note that
Hence
Now, from Theorem 11, there exists a constant
K such that
. Then,
Consequently, by (
30), (
32), and (
33) we have that