Abstract

We show conditions for the existence, continuity, and differentiability of functions defined by , where is a function of bounded variation on with .

1. Introduction

Let be a complex function defined on a certain subset of . Many functions on functional analysis are integrals of the following form:

We discuss the above function , where the integral that we use is that of Henstock-Kurzweil. This integral introduced independently by Kurzweil and Henstock in 1957-58 encompasses the Riemann and Lebesgue integrals, as well as the Riemann and Lebesgue improper integrals.

In Lebesgue theory, there are well-known results about the existence, continuity, and differentiability of . For Henstock-Kurzweil integrals also there are results about this, for example, Theorems 12.12 and 12.13 of [1]. However, they all need the stronger condition: is bounded by a Henstock-Kurzweil integrable function . We provide other conditions for the existence, continuity, and differentiability of .

2. Preliminaries

Let us begin by recalling the definition of Henstock-Kurzweil integral. For finite intervals in it is defined in the following way.

Definition 2.1. Let be a function. One can say that is Henstock-Kurzweil (shortly, HK-) integrable, if there exists such that, for each , there is a function (named a gauge) with the property that for any -fine partition of (i.e., for each , ), one has
The number is the integral of over and it is denoted as .
In the unbounded case, the Henstock-Kurzweil integral is defined as follows.

Definition 2.2. Given a gauge function , one can say that a tagged partition of is -fine, if (a)a = ,,(b) for all,(c) .

Definition 2.3. A function is Henstock-Kurzweil integrable on , if there exists such that, for each , there is a gauge for which (2.1) is satisfied for all tagged partition which is -fine according to Definition 2.2.
Let be a function defined on an infinite interval , One can suppose that is defined on assuming that . Thus, is Henstock-Kurzweil integrable on if extended on is HK-integrable. For functions defined over intervals and One can makes similar considerations.
Let be a finite or infinite interval. The space of all Henstock-Kurzweil integrable functions over is denoted by . This space will be considered with the Alexiewicz seminorm, which it is defined as follows: where the supremum is being taken over all intervals contained in .

Definition 2.4. Let be a function, where is a finite interval. The variation of over the interval is defined as follows:
We say that the function is of bounded variation on if . Now if is a function defined on an infinite interval , then is of bounded variation on , if is of bounded variation on each finite subinterval of and there is such that for all . The variation of on is .
Given an interval , the space of all bounded variation functions on is denoted by . We set . The following are some classical theorems that are used throughout this paper. The first is given in [2, Lemma 24] and is an immediate consequence of [1, Theorem and Corollary ].

Theorem 2.5. If is a HK-integrable function on and is a function of bounded variation on , then is HK-integrable on and

Theorem 2.6 ([1] Chartier-Dirichlet's test). Let and be functions defined on . Suppose that(i) for every , and defined by is bounded on ; (ii) is of bounded variation on and .Then .

Definition 2.7 (see [3]). Let . A function is on , if for every , there exist and a gauge on such that whenever is a -fine subpartition of (i.e., is -fine and the tags belong to ) and
We say that is on , if can be written as a countable union of sets on each of which the function is .
If is a function on , then we use the notation for the partial derivative of with respect to the second component .

Theorem 2.8 (4, Theorem). Let . If is such that (i) is on for almost all ; (ii) is HK-integrable on for all .
Then is on and for almost all , if and only if, for all . In particular, when is continuous at .

3. Main Results

All results in this paper are based on functions in the vector space . Note that , where is the space of Lebesgue integrable functions. Indeed, the function defined by is in . However, for bounded intervals , functions in are Lebesgue integrables on .

To facilitate the statement of these results, it seems appropriate to introduce some additional terminology. If is a function and , we say that satisfies Hypothesis (H) relative to if () there exist and , such that, if then for all .

This type of condition plays a major role in the results of the present work.

Theorem 3.1. Let and   be functions. If , and satisfies Hypothesis (H) relative to , then exists for all in a neighborhood of .

Proof. It follows by Theorem 2.6.

Theorem 3.2. Let and be functions such that (i), is bounded, and (ii) is continuous for all . If satisfies Hypothesis (H) relative to , then the function defined in Theorem 3.1 is continuous at .

Proof. There exist and , such that, if then for all . From Theorem 3.1, exists for all .
Let be given. By Hake's Theorem, there exists such that for all . On the other hand, as there is such that for each ,
Let . From Theorem 2.5, it follows that for every and every , where the second inequality is true due to (3.4). This implies, since , that
Analogously we have that
Therefore, for each ,
By hypothesis, is Lebesgue integrable on , is bounded, and is continuous for all . From this it is easy to see, for example using [1, Theorem 12.12], that defined as is continuous at . This implies that there is such that for every ,
Let . Then for all ,
Thus, from (3.5), (3.11), and (3.13), , for all .

Theorem 3.3. Let . If and are functions such that (i), is measurable, bounded, and (ii) for all , satisfies hypothesis (H) relative to , then

Proof. From condition (ii) and by the compactness of , we claim that there exists such that, for each , , for all .
For each and , let . Observe, by Theorem 2.5, for all .
So, for each , is HK-integrable on and is bounded for a fixed constant. Moreover, by Theorem 3.1 and Hake's Theorem, for all .
Therefore, by dominated convergence theorem, is HK-integrable on and
Now, since is Lebesgue integrable on , is measurable and bounded; it follows by Fubini's Theorem that
Consequently,
So by Hake's Theorem,

Theorem 3.4. Consider and functions, where and the partial derivative exists on and is bounded and continuous. If such that(i)there is for which for all , and (ii) satisfies Hypothesis (H) relative to ;then is differentiable at , and

Proof. It is not difficult to prove, using conditions (i), (ii), and the Mean Value Theorem, that there exist and such that, for each , for all .
Consider with . In order to show (3.22), we use Theorem 2.8. The function is differentiable on for each , so is on for all . Also, by (3.23) and Theorem 2.6, is HK-integrable on for all . Then if is continuous at , and for all . The first affirmation is true by (3.23) and Theorem 3.2. The second affirmation is true due to (3.23) and Theorem 3.3.

Remark 3.5. In the previous theorems the kernel satisfies , for all . Moreover, if will satisfy for all , then, when (a version of Riemann-Lebesgue Lemma).

4. Applications

If , then its Fourier transform at is defined as follows:

Talvila in [2] has done an extensive work about the Fourier transform using the Henstock-Kurzweil integral: existence, continuity, inversion theorems and so forth. Nevertheless, there are some omissions in those results that use [2, Lemma 25(a)]. Also Mendoza Torres et al. in [5] have studied existence, continuity, and Riemann-Lebesgue Lemma about the Fourier transform of functions belonging to . Following the line of [5], in Theorem 4.2, we include some results from them as consequences of theorems in the above section.

Let and be real-valued functions on . The convolution of and is the function defined by for all such that the integral exists. Various conditions can be imposed on and to guarantee that is defined on , for example, if is HK-integrable and is of bounded variation.

Lemma 4.1. If , then .

Proof. Since is of bounded variation on , then and exist. Suppose that . If , there exists such that , for all . If , there is such that , for all . This shows that or , which contradicts , so . Using a similar argument, we show that .

As consequence of Lemma 4.1, the vector space is contained in . So the next theorem is an immediate consequence of the above section.

Theorem 4.2. If , then (a)exists on . (b) is continuous on .(c). (d) Define and suppose that , then is differentiable on , and (e) If , then for all .

Proof. First observe that for all . Then, each satisfies Hypothesis (H) relative to .
Theorem 3.1 implies that exists for all and, since , exists. Thus, exists on .
From Theorem 3.2, is continuous at , for all .
It follows by Remark 3.5 and (4.4).
It follows by Theorem 2.8 in a similar way to the proof of Theorem 3.4.
Take and let . Then, for each and all ,
Thus, for every , satisfies Hypothesis (H) relative to . Now, observe that and is measurable and bounded. So by Theorem 3.3, for all .
On the other hand,
Thus, since , dominated convergence theorem implies that but from (4.6), we have that Therefore, by Hake's Theorem,

If , then its Laplace transform at is defined as follows: Here, also the Laplace transform is considered as Henstock-Kurzweil integral.

Theorem 4.3. If , then (a) exists for all with . (b) If , then is continuous on for all , and is continuous on for all .

Proof. It is an easy consequence from Theorems 3.1 and 3.2, since for all , with .

Moreover, the Riemann-Lebesgue Lemma holds the following.

Theorem 4.4. If and , with , then

Proof. It results by Remark 3.5 and (4.4), because is in .