1. Introduction
The problem of equivalence between differential and integral problems is fundamental to the use of operator methods to study solutions to such problems. A suitable integral operator allows us to prove the existence of fixed points and thus solutions to differential problems. The problem depends on how we define the solution or what we assume about the functions in the problem. It is not well studied for vector-valued functions with values in Banach spaces, and it is of particular interest to consider the case of weak topology (cf. [
1,
2]), which is different from that of real functions. The brief discussion of the case of strong derivatives, and why we study the weak topology here, is presented later, in Remark 1.
The use of methods based on the properties of integral operators to solve given differential problems always requires studies of the equivalence of the problems. As we show in this paper, especially for operators of fractional order, this is an absolutely fundamental step, without which most results about the existence or regularity of solutions will not hold. In practice, most papers on this subject start by presenting an integral problem equivalent to the initial differential problem, and this step is not always supported by correct reasoning, e.g., omitting the problem of the function space in which the problem is studied. Integral operators are much more convenient to study, and our results will provide a basis for their use in the study of fractional order differential problems. Our paper will verify and even extend such studies (for vector functions, among others).
The problem we are discussing is absolutely fundamental to the study of differential equations, including, of course, those of integer order. Surprisingly, the equivalence problem for vector-valued functions has so far been studied in the case of integrals and strong derivatives [
1] of integer order. But, even in this base case we have a variety of different integrals and solution spaces. Basic results for weak solutions and pseudo-solutions are also collected in the same paper. In the cases considered there, however, the integral operator was only a Volterra operator. In the study of fractional order problems, we study weakly singular operators, so the problem has to be studied anew. We fill this gap in our paper.
The basic cases of the Riemann–Liouville integral and the Caputo derivative for real-valued functions are quite well studied [
3,
4]. However, many models based on other fractional order integrals (e.g.,
-Hilfer integrals) remain to be studied. This problem is often overlooked or neglected by authors. And it requires both the identification or construction of corresponding differential and integral operators, and it also depends strongly on the function spaces considered on which we study these operators (and thus the solution spaces). Of course, the importance of checking the equivalence of differential and integral problems has been recognized many times. An excellent example is the article [
5], where previous results on fractional order equations with impulse effects were reviewed and improved (with regulated solutions). In this paper, however, we will concentrate on problems for vector functions and weak derivatives. It is important to emphasize the role and usefulness of the study of conditions expressed in weak topology for practical issues. In this context, let us choose some examples. Applications of this topology include the study of random and stochastic equations [
6,
7], dynamical systems [
8,
9], or differential inclusions (cf. [
10] for the case of fractional derivatives).
These studies are usually carried out on the most commonly defined fractional order integral so that all the previous ones are included in a single result. Thus, we will deal here with the most commonly used and very general notion of Stieltjes-type integrals, i.e., the
-Hilfer integral with respect to the
function. The problem has recently been studied in [
4] for real-valued functions. Note that the case of differential problems for vector functions with strong derivatives is similar to that for real functions and does not present significant mathematical difficulties for this theory. However, considering the interesting problems with weak differentiability of vector functions, we will concentrate on the still unsolved problem for integrable functions in the Pettis sense (and thus weak). We should note here that in the considered case we cannot expect the equivalence of weak-type problems in arbitrary space (cf. [
11], where the lack of equivalence in a general case is proved). This is another reason to study the case of weak solutions and pseudo-solutions. Here, the foundations of the problem are given in [
12,
13] but only for some particular fractional operators. The case under consideration is fundamentally different from that of real functions since it is necessary to ensure that weak-type derivatives [
1,
14,
15,
16] are appropriate. For a description of this problem, see [
17] or for fractional derivatives, see [
13].
In this paper, we investigate solutions in the weak sense for fractional-order differential equations for a very general class of fractional operators and taking into account the natural requirement of Hölder-type continuity of solutions. As previously claimed, many fractional order derivatives and integrals exist. The aim of this paper is to obtain results for a wide range of them. This includes the definition of inverse differential and integral operators that go far beyond the classical cases of the Riemann–Liouville integral and the Caputo derivative. There is also the problem with vector-valued functions and weak topology. How is the derivative supposed to exist? This has been known since the study of integer order equations (see [
1]) but is particularly relevant for fractional-order equations.
In the context of differential equations, the possibilities of using the Pettis integral were already noted in the papers [
2,
18]. These results were initially unappreciated, but after a systematic study of the relationship between weak-type solutions and Pettis integrals [
1], research accelerated. This is a very different case from strong derivatives. Also of interest was the question of how to extend and unify ordinary differential equations with equations of fractional order. Over the years, more and more attention was paid to the theory of Pettis integrals, and there was a big step between the unification of fractional calculus and the theory of Pettis integrals (see [
11] and references therein).
In particular, it was proved in [
13] that even if the right-hand side of a pseudo/weak Caputo fractional order problem is weakly continuous, the equivalence between the differential and integral forms can be lost. It thus follows that if
(resp.
denotes the
-tempered Riemann–Liouville (resp. Caputo) fractional pseudo-derivative
, and with parameter
, then there exists
such that
It is evident that this is also applicable in the case of tempered Hilfer fractional pseudo/weak derivatives.
Since the aim of this paper is to demonstrate the equivalence of differential and integral problems of fractional order with Pettis integrals, we should also pay attention to applications. This concerns the study of differential equations of fractional order with given initial or boundary conditions in the case of vector functions. Such studies have been carried out, among others, for the weakly singular Riemann–Liouville fractional Pettis integral, for pseudo-solutions of fractional BVP’s, and for the existence of random solutions for a class of partial random integral equations using Hadamard fractional integral, under the assumption of Pettis’s integrability. A further important step is the paper [
12] and its consequences, i.e., [
13].
In order not to prolong the introduction of this paper, we will place the discussion of some other results that are relevant to the results obtained here, and the comparison with previous results of this type, in
Section 7.
Finally, it is worth mentioning here a paper in which the Pettis integral is combined with another operator of fractional order, namely the Hadamard operator. The paper [
12] proves the existence of fractional point solutions of differential equations of the Caputo–Hadamard type of the Pettis integral. This is already a step towards the general integral operators of fractional order studied in this paper. With the introduction of the
-Hilfer fractional derivative, containing a wide class of particular cases from the choice of
, the parameters
a,
b and the limits
or
. This same property is also valid for the
-Riemann–Liouville fractional integral. In the particular choice of the function
, with
, we obtain the version of the Hadamard–Pettis fractional integral discussed, for example, in [
12]. In order to maximize the generality of the results, this paper considers an equivalence problem for such a general class of fractional integrals and and corresponding pseudo-derivatives of fractional order.
It should be noted that the previous results concerned the special case of weak integrals of fractional order and a rather narrow class of E target spaces. Our results allow us to extend them considerably.
2. Preliminaries
Let E denote a real Banach space with a norm and its dual. Let denote the space E when it is endowed with the weak topology (i.e., generated by the continuous linear functionals on E). Moreover, let denote the Banach space of continuous functions from I to E, with its weak topology. A function is said to be weakly measurable on I if for every the real function is Lebesgue measurable on I. Denote by I the compact interval .
We need the following definitions [
19] in this paper:
Definition 1. A weakly measurable function is said to be Pettis integrable on I if
- 1.
x is Dunford integrable on I, that is, for every ;
- 2.
For any measurable , there exists an element in E denoted by such that
With
, we denote the space of
—valued Pettis-integrable functions on the interval
I. Immediately, it should be recalled that this space is normed but is not a complete space. Thus, we cannot reduce the case studied here to the results concerning Banach spaces (cf. [
1]). Instead, this paper explores some subspaces of this function space. These subspaces consist of weakly Hölder continuous functions.
The problem of differentiability of weak integrals is more complicated than the strong differentiability. Let us recall, that there exists a strongly measurable and Pettis-integrable function
such that the indefinite Pettis integral
is not weakly differentiable on a set of positive Lebesgue measures or even nowhere weakly differentiable [
17]. And this is one of the difficulties in achieving equivalence of differential and integral problems in weak topology.
The solution—although, as we shall see, not entirely sufficient—is to adopt a different, much more general definition of derivative [
2].
Definition 2. A function is said to be pseudo-differentiable if there exists a function such that, for every , the real-valued function is differentiable a.e., to the value . In this case, the function y is called the pseudo-derivative of the function x and will be denoted by .
If the null set is independent on , then x is said to be almost everywhere weakly differentiable on I and y (in this case) is called weak derivative of x (denoted by ) and exists almost everywhere on I. In particular, when , it is clear that the pseudo- and almost everywhere weak derivatives coincide with the classical derivatives of real-valued functions.
If the space E has a total dual . (i.e., the space contains a countable subset separating the points of E), then the pseudo-derivative, if any, is uniquely determined up to a set of measure zero. Note that the weak absolute continuity of Banach-valued functions does not necessarily imply strong or everywhere weak differentiability. This is another argument for the need to study the equivalence of fractional order derivatives problems beyond the superpositions of absolutely continuous functions.
Let
denote the pseudo differential operator (resp.
for the weak one). Similarly, for any
being a positive increasing real-valued function such that
, for all
, we define some weak-type fractional derivative as follows
Remember that the absolutely continuous real-valued function is a.e. differentiable. This is one of the few properties of the real-valued function that does not carry over to arbitrary Banach spaces: In fact, even in separable spaces, there is a weakly absolutely continuous function which has no pseudo-derivative [
14]. We also note that a function can be pseudo-differentiable and at the same time neither weakly nor strongly differentiable. Let us first recall the relationship between the Pettis integral and the corresponding definition of differentiability.
As claimed above, with the notion of pseudo-derivative, we are able to make a step towards the equivalence of integral and differential (not yet fractional) problems:
Proposition 1 ([
16,
19], Theorem VIII.3 in [
14])
. A function is an indefinite Pettis integral if and only if f is weakly absolutely continuous having a pseudo-derivative on I. In this case, and The following result can be obtained by assuming weak absolute continuity of the function:
Proposition 2 ([
20] and Theorem 5.1 in [
21])
. The indefinite integral of Pettis-integrable (resp. weakly continuous) function is wAC and it is pseudo- (resp. weakly) differentiable with respect to the right endpoint of the integration interval and its pseudo- (resp. weak) derivative equals the integrand at that point. 3. Fractional Integrals and Corresponding Function Spaces
We now turn to the case of derivatives and fractional integrals, and this idea of multiplication by a real and bounded function under the sign of the Pettis integral was already considered in the original paper [
19]. We will now use this to define a very general fractional integral of a vector function.
What is new here? Compared to [
19], here we have to consider operators on completely different function spaces (consisting of continuous functions), which will allow us to study the equivalence of differential (solutions are continuous or weakly continuous) and integral problems (i.e. the values of the integral operator should also be in such spaces).
Definition 3. Let be a positive increasing function such that , for all . Let be vector-valued, weakly measurable function. We define
- 1.
(ψ-tempered Riemann–Liouville fractional Pettis integral:) The tempered Riemann–Liouville ψ-fractional Pettis integral of f of order and with parameter is defined by
where . For completeness, we define . - 2.
(ψ-tempered Riemann–Liouville fractional derivative:) The ψ-tempered Riemann–Liouville fractional weak (resp. pseudo) derivative of order , , and with parameter is defined as
- 3.
ψ-tempered Caputo fractional derivative: The ψ-tempered Caputo fractional weak (resp. pseudo-) derivative of order , , and with parameter is defined as
For any continuous
with a positive continuous derivative
on
I (cf. Corollary 3.41 in [
19]) ensures that
exists for every
and by Proposition 2, it is easy to see that
Some remarks on the choice of integral and differential fractional operators. In recent years, there has been a tremendous growth both in the definition of fractional derivatives and in their applications. In short, the
-Hilfer operator discussed in this paper has some properties that are important for our problem. The generalized fractional calculus considered here unifies several existing ones, especially those considered in differential equations used to model some phenomena in physics. Although these conclusions are well known, we will briefly summarize them for the convenience of the reader. In particular, as special cases, certain choices of
yield several classical models of fractional calculus (see [
4,
22]).
It also seems useful to compare studies of these operators in the case of norm topology. Originally they were studied in certain weight spaces of continuous or absolutely continuous functions [
23], and only later the need to preserve the regularity of the values of integral operators by Hölder spaces was recognized (e.g., [
4]). This includes several cases that have been studied separately.
In the case and , it is a classical fractional calculus with Riemann–Liouville and Caputo derivatives. There is a very large literature on this subject, so we will not go into the details.
TIn the case of and , we obtain the tempered fractional calculus, which has been studied in recent years because of its applications, for example, in stochastic and dynamical systems. In this context of tempered calculus, fractional diffusion applies an exponential tempering factor to the particle jump density.
If and , with we obtain the Hadamard calculus (the generalized version of the classical Hadamard calculus provided that ). It is useful in the context of the study of the mean first-passage time as a function of in order to highlight the difference with respect to the time-fractional model considered for the anomalous diffusion. The function g controls the anomalous diffusion at intermediate times.
Next, we consider the case
(
) and
. We obtain the Katugampola fractional integral operators (see also the case the Erdélyi–Kober fractional operators. When
and
, there are again some applications in the description of the anomalous subdiffusion [
24,
25].
In order not to focus the reader’s attention too much on this issue, let us simply note that the ability to select the -function allows many real world phenomena to be modeled by its selection. Let us mention only to describe subdiffusion in a medium with a time-evolving structure, or in the study of the relation characteristic of superdiffusion processes.
However, finding and studying the case of weak topology leads to further complications.
Remark 1. Instead, it is essential to keep in mind the differences between weak and strong topology, which we believe provides a strong basis for the separate research proposed in this paper. The case of the norm topology on E, as metrizable, is similar to the study for real functions with the replacement of the Lebesgue integral by the Bochner integral. If the functions under consideration are continuous in the sense of such a topology [26], the results obtained for real functions are similar to this case, the derivatives being in the strong sense [4]. In this paper, we will therefore concentrate on the much more complicated case of weak topology (as claimed in [27,28], for instance). Since the indefinite Pettis integral of a function does not enjoy the strong property of being a.e. weakly differentiable (see [17]), then does not necessarily hold for arbitrary . The formula is not uniquely determined unless E has total dual . Evidently, according to (e.g., [16], p. 2), it may happen that with g being weakly equivalent to f (but they need not be necessarily almost everywhere equal).
We are now going to look at what the inverse operators of the fractional integral operators we are studying are.
Let
be a real-valued positive increasing function such that
, for all
,
and
be a vector-valued function with a pseudo-derivative in
. Then, by Proposition 1, we obtain the result for pseudo-derivatives:
In particular, if
,
For the convenience of the reader, the following function spaces are provided:
Definition 4 ( [
19])
. Let denote the space of E-valued Pettis-integrable functions on I. As its subspace, let us consider First, let us note some limitations on the consideration of fractional order integrals of vector functions. We are already aware of some results which indicate the need for in-depth studies of the equivalence problem between differential and integral equations. In our considerations, we need to go beyond the results of Pettis [
19] (and [
33] in the case of fractional operators for real-valued functions) and use the following results:
Proposition 3 ([
19])
. A function if and only if for every , with . Remark 2. As known from our previous research (e.g., [4,11]), exists (is convergent) if . This is true for any if E is a reflexive space. It should be added, however, that the Pettis integrability conditions have not yet been fully explored, even in the case of . A partial characterization and acting conditions are given in the following: Proposition 4 (Lemma 2 in [
12])
. Let be a positive increasing function such that , for all . For any and , then- 1.
For any , we have
- 2.
For any , we have on
The next step is to specify the appropriate function space as the domain of the operators. As in the case of real functions (cf. [
4]), a variant of the Hölder condition plays an important role, as further demonstrated by Hardy and Littlewood in [
33]. By reasoning as in [
13], we know that there is
such that
is “meaningless" even for some Hölder continuous functions.
This points out that the equivalence between Caputo fractional differential and corresponding integral problems is no longer necessarily true outside the space of the absolutely continuous (or weakly absolutely continuous). In order to avoid such an equivalence problem, we proceed in a different way by showing the equivalence using the following definition
Definition 5. Let be a real-valued positive increasing function such that for all and is a weakly measurable vector-valued function. For any , we define a Caputo-type fractional derivative Let us introduce a subspace of the space of Pettis-integrable functions. To do it this, we need to recall Hölder spaces.
For
, the Hölder space
consists of all functions
such that
and for
, there exists a constant
(independent on
) for which
. Obviously,
if the following seminorms
is finite. Define also its vector-valued counterpart:
and
where
, and 0 denotes the zero element of
E.
However, the equivalence problem is still not solved: recall that the weak absolute continuity of is a necessary (but not sufficient) condition for the existence of a pseudo-derivative of . Moreover, would exist on I, and then necessarily with
Obviously, , and the equality holds when, e.g.,
Since the weak continuity of a function
f implies strong measurability ([
20], p. 73) and
, it obvious
. Also, for any
we know that (cf. Remark 2)
exists and by definition of the Pettis integral, for any
, we have
Spaces of the type we are considering (functions that satisfy the condition ) are sometimes called little Hölder spaces. The Remark below shows why we are interested in these spaces.
Remark 3. Let us remark that (see [34], Exercise 86(b), p. 353, [35], Lemma 3.1) for any we have Therefore, for any , Hence, for any we have (in view of Proposition 4) Fractional integrability in the relevant sense of pseudo-derivatives will now be examined.
Lemma 1. Let and , weakly absolutely continuous with a pseudo-derivative be uniquely determined up to a negligible set. Then the Stieltjes integral exists and the following integration by the parts formula holds true: Proof. Since
(cf. Theorem 3.4 in [
19]) and
(cf. Corollary 3.41 in [
19]), then by the definition of the Pettis integral on
, there exists an element in
E denoted by
such that for any
By the definition of the pseudo-derivative of
f on
we have for any
. Consequently, since
(namely,
), we obtain
□
The next lemma gives a condition that ensures that the two integrals and coincide in the space of weakly absolutely continuous functions:
Lemma 2. Let be a positive increasing function such that , for all , and , which is weakly absolutely continuous with a pseudo-derivative uniquely determined up to a negligible set. Then, exists and In particular, if E is a reflexive space, it holds for all .
Proof. Let
with a pseudo-derivative in
,
. Then
having a pseudo-derivative in
,
as well. Since
, it follows (cf. Remark 2) that
exists and by definition of Pettis integral, for any
, we have
By Proposition 1, we have
for
. Therefore,
Since the indefinite integral of
is weakly absolutely continuous and has a pseudo-derivative on
I, it follows from (
9) that
Keep in mind the following:
for
, we obtain
Consequently, according to Proposition 4
So
is weakly absolutely continuous and possesses a Pettis-integrable pseudo-derivative on
I. By Proposition 1
Obviously, , by virtue of Remark 3. Moreover, by Remark 2, this is also true for all when E is reflexive. □
Remark 4. In Lemma 2, the pseudo-derivative of is, in fact, “an almost everywhere weak derivative” on I: To see this, we notice that . Hence, by Proposition 4, . By (10), we have Recall [30] that the indefinite Pettis integral of a weakly continuous function is weakly absolutely continuous and possesses a weak derivative on I. This is in the view that the null set (where the derivative of fails) is invariant for all , yields the almost everywhere weak differentiability of on I. And now one of the key results of the paper for vector functions, which is an analogue of the Hardy–Littlewood theorem [
33] and the result of [
4] for real-valued functions. Despite all the differences between integrals and topologies, fractional order operators preserve very well the Hölder orders of the functions on which they act.
Theorem 1. Let be a real-valued positive increasing function such that , for all , , . Then, the vector-valued integral operator has values in , i.e.,for any , and moreover,where . Proof. Let
,
; it follows that
exists (cf. Remark 2) and, by definition of the Pettis integral, for any
, we have
. Now let
. We have the following estimate
Hence,
maps
into
and for any
,
□
We are now going to show an analogous result for .
Since
for every
for
and since
We can repeat our argument with slight modifications. Namely, if one argues as in the proof of Theorem 1, then for all
and
, we have
Since
we obtain
In view of the mean value theorem,
and then
Since
we obtain the following
Thus, in the vector-valued case we have also shows interesting results for :
Theorem 2. Let be a positive increasing function such that , for all , . Then for any It is worth noting that the theorem proved is a far-reaching extension of the classical Hardy–Littlewood (Theorem 4 in [
33]) result. In the case
and
, we have:
Corollary 1 ( Theorem 4 in [
33])
. Suppose that , where , , . Then the Riemann–Liouville fractional integral operator has values in , , andwhere is a function of p and α only. In the above result, the Riemann–Liouville integral operator still acts between Lebesgue spaces. In the context of Hölder continuity of images of -integrable functions, we also extend the following:
Corollary 2 (Proposition 3.2(3) in [
3])
. For or and , the fractional Riemann–Liouville integral operator is bounded from into a Hölder space , and hence, for , is Hölder continuous with exponent . It remains to examine the operator once we have restricted its domain to the appropriate Hölder space.
Proposition 5. Let be a positive increasing function such that , for all , . Then Also, there exists a constant such that Proof. Let
for any
(in particular for any
). It follows (cf. Remark 2) that
exists and by definition of the Pettis integral, and for any
, we have
Also, by (
8), we have
. Obviously, since
, then automatically
or
. Since
it follows
Therefore, the result follows from Theorem 2. □
Finally, we proved a vector-valued extension for the case of generalized fractional integral operators acting between Hölder spaces:
Corollary 3 ( Theorem 1 in [
4])
. Let be a positive increasing function such that for all with . For arbitrary , such that , we haveIn particular, for any with the operator maps the Hölder space into itself.
Theorem 3. Let be a positive increasing function such that , for all , such that . Then Also, there exists a constant such that As we have already noted, it is possible to consider the norm topology on E in a way that is similar to the case of the weak topology (but not vice versa). So, we have:
Corollary 4. Let be a positive increasing function such that , for all , such that . Then, Also, there exists such that Proof. We will give a sketch of the proof showing how to obtain results for this case using our main theorem. Let
. Since
we obtain
. Also, any
we know that
Without loss of generality, assume that
. Then, there exists (as an immediate consequence of the Hahn-Banach theorem)
with
such that (cf. Theorem 3)
Note that the function
is chosen with fixed
t and
h, i.e., to obtain a universal estimate for the Hölder constant we must estimate these expressions independently of these quantities. However, by definition, we have estimated
That explains how this came about. □
As a result, we obtain the following
Corollary 5. Let be a positive increasing function such that , for all , and . Then, we haveIn particular, Proof. Choose
. Since
it follows from Theorem 3,
Since
we obtain
□
6. Application
Next, we will prove the following for all without imposing an absolute continuity condition on f.
Let
and
. Since
for any
, Proposition 2 can be combined with Proposition 4 to ensure (in view of Theorem 6) that
Also, for any
we have
Let
be the separable space of all real-valued sequences that converge to 0. The topological dual
is known to be isometrically isomorphic to
(the space of all summable real-valued sequences). Define
by
It is immediate that
acts from
into
and is weakly measurable (even strongly measurable by the Pettis measurability theorem [
19]). We recall that there exists a unique
corresponding to each
such that for any
such that
we have
Of course, the estimate
yields (in view of
) that
for any
. According to [
13],
has the Riemann–Liouville fractional pseudo-derivative of any order
, while the pseudo-derivative does not exist on
. Furthermore, Theorem 6 gives
.
Example 1. Define by (with ). Consider the followingcombined with appropriate initial or boundary conditions. Of course, (22) and the corresponding “standard" integral formare not (in general) equivalent outside the space of absolutely continuous functions. Even the (restrictive) assumption that is not sufficient for the equivalence between (22) and (23): Obviously,Since does not exist, x given by (23) is not a solution of (22). On the other hand, in view of (19), we can convert (22) into the following integral form: In view of (18), the equivalence between (22) and (27) holds and we simply arrive at (22). Remark 5. Recall that Theorem 6 (hence (18) is true for any Banach space E even if E has no total dual. In such a case, any two pseudo-derivatives of a function are weakly equivalent on I. Suppose exists and are two pseudo-derivatives of ; then, for every there exists a null set such thatfor . Hence, Consequently, the results of Theorem 6 and (18) are independent of the choice of a pseudo-derivative of . In the following, we construct an example that shows this: The equivalence between the Caputo-type differential problems and the corresponding “standard” integral form may be lost for the functions with values in the Banach spaces without total dual. However (given (
18), the equivalence holds in arbitrary spaces.
The following example is also a warning against too quickly attempting to take equivalence results for real functions as true for results for vector-valued functions.
Let
be the Banach space (having no total dual) of bounded real-valued functions on
and let
be of positive measure, and define
by
We know that [
19], Example 9.1,
with
for each
. So, it is quite obvious that
for any
Example 2. Consider the problem (22) with f defined by (25). The corresponding “standard" integral form isfor . According to Example 9.1 in [19], the function has two pseudo-derivatives 0 and f on , that differ on the set of positive measure . Therefore, eitheror Consequently, we conclude that on J we have .
On the other hand, given (19), we can transform (22) into the following integral form: Since for arbitrary it follows, in view of (18) (bearing in mind Remark 5), the equivalence between (22) and (26) holds, and we simply arrive at (22). Namely, for any , we have Remark 6. Recall that due to Theorem 2 in [13], for any infinite dimensional Banach space E that fails cotype, there is a function and a weakly absolutely continuous, nowhere weakly differentiable function g defined on such that We also know that , by reasoning as in proof of Theorem 1 in [13]. Consider now the problem (22) with , with . Recall that This yields is weakly absolutely continuous and nowhere weakly differentiable on . So, the problem (22) with Caputo-type fractional weak derivatives would make no sense. In many papers, authors will not consider the Hölder continuity. They will look for solutions of differential or integral equations only in spaces of continuous functions. In such a case, the following conclusion is useful:
Corollary 6. The operator , for , is weakly compact.
Proof. Let be bounded subset of . Let be a sequence in convergent weakly to f in . From Theorem 5, it follows that .
Since
is weakly convergent in
, it is bounded, weakly equicontinuous, and weakly pointwisely convergent. Again, by applying Theorem 5, we obtain the same property for
(with the estimation of of the modulus of the weak continuity by
), we conclude that the latter is also weakly convergent in
. Indeed, for each
and
we obtain
Moreover, for any
and
, we have
Hence, in view of (
16)
it follows that
Hence, the weak sequential equicontinuity of the image of
under
follows. Also, in view of
, there exists
such that
the uniform boundedness also follows.
Thus,
is weakly sequentially continuous and the set
is weakly sequentially compact in
. By the Eberlein–Šmulan thorem (e.g. [
36]), it is also weakly compact in
. Therefore,
maps bounded sets in
into weakly compact sets in
. Thus,
is weakly compact. □
7. Comparison with Previous Results: Final Comments
The problem of the equivalence of differential and integral problems studied in this paper has a long history. Our results directly extend those of fractional order operators and are applied to the study of fractional order differential equations. For generalized fractional order integral operators, i.e., differential problems with derivatives with respect to another function (with some special choice of such functions), this applies to the case of real-valued functions (e.g., [
3,
4,
35]) for the case of vector-valued functions with strong topology (e.g., [
26]) and the basis for vector-valued functions with derivatives considered in the weak sense or pseudo-derivatives in [
11,
37]. We will now compare the more relevant of these (and other) papers with the results obtained.
As we have already mentioned, it is now possible to extend the known results, among others mentioned in the Introduction, to the case of pseudo-solutions, i.e., for -Hilfer integrals of fractional order in the integral form of the considered problem. However, this can be performed directly by repeating the proofs of other papers, inserting only our results on equivalence problems, and would not show the relevance of our theorems.
In order to focus the readers’ attention on the main objectives and results of this paper, we will present only some extensions of the already known results, choosing those most closely related to fractional order operators and derivatives in weak topology. We will not focus on correcting errors in previous papers (as in the paper of [
5]) but will point out the possibility of considering the issues under study with much more general derivatives and integrals of fractional type, which, in addition to obtaining generalizations of previous results, allows them to be unified. The derivatives and integrals considered, as we have already mentioned, include many previously studied operators.
We will limit ourselves mainly to a few of the papers cited in the Introduction and to the most advanced ones.
First, let us recall the result from [
38] about the existence of ‘weak-mild’ solutions for the problem
with boundary-value condition
in reflexive spaces.
Using the approach from our paper, first of all, the solution definitions should be improved to refer to pseudo-derivatives (which is not there), cf. Proposition 1. According to the results of our paper, to improve the assumptions that guarantee that the solution operator is well defined (cf. Theorem 1). Furthermore, the problem can be considered with the fractional derivatives and the integrals as in Definition 3. Of course, we also obtain greater regularity of solutions (not necessarily in spaces of absolutely continuous functions). This paper can be corrected and extended significantly.
In the next step, the research was extended to the existence of pseudo-solutions and weak solutions for nonlocal boundary problems using Riemann–Liouville operators. This required a complete study of the Pettis integral of fractional order and the corresponding spaces of integral functions in the Pettis sense.
This research showed that the close relationship of the spaces associated with fractional-order operators allows for weaker assumptions that guarantee a good definition of the operators. It was also an indication that such results could and should be extended to a larger class of fractional-order operators and their greater regularity obtained. Further important steps in this direction can be found in the papers [
4,
11,
12,
13].
The situation is similar in the paper [
37], where, among others, the existence of solutions to the fractional differential problem given by
where
is a fractional pseudo-derivative of a weakly absolutely continuous and pseudo-differentiable function
is established. The results go beyond reflexive spaces (cf. [
39]) and are carefully proved. These proofs can be carried out without major changes for the case proposed in this paper, i.e., with the inclusion of much more general fractional order operators (cf. again Definition 3). This allows them not to be proved separately for, e.g., the Hadamard operator. It is worth mentioning also a paper in which the Pettis integral is combined with the Hadamard operator. In the paper [
12], the existence of solutions of differential equations of the Caputo-Hadamard type of the Pettis integral is proved. This was a step towards the general integral operators of fractional order studied in this paper (cf. [
40]). Similarly, the multi-term fractional differential equation studied there can be generalized using our treatment of the topic. Again, we can consider Hölder spaces and solutions more regularly.
In order not to overstretch the concluding remarks, we will only point out that the direction discussed in the paper to study the regularity of solutions of fractional order equations in Hölder-type spaces is currently being pursued for the case of real functions (and thus without problems related to weak topology). Let us mention a recent paper [
4].